Araujo Et Al - 2000 2
Araujo Et Al - 2000 2
Chapter 26
L. M. Araújo J. R. Cerqueira
Petrobrás E & P/GEREX/GESIP L. C. da S. Freitas
Rio de Janeiro, Brazil
Petrobrás/Cenpes/DIVEX/CEGEQ
Rio de Janeiro, Brazil
J. A. Trigüis
UENF
Macaé, Brazil
Abstract
In the Permian petroleum system of the Paraná Basin, organic-rich source rocks (Assistência Member
of the Irati Formation) were thermally matured by the heat of Cretaceous basic igneous intrusions (dia-
base sills) of varying thicknesses (few centimeters up to 240 m). This constitutes an atypical petroleum sys-
tem, characterized by the synchronism of generation and migration processes with the magmatism
(138–127 Ma). During this period, countless generation and migration pulses occurred from multiple gen-
eration kitchens that were not related to the basin depocenter, in sharp contrast to basins in which thermal
maturation is controlled by increasing burial.
The total volume of hydrocarbons generated by the thermal effect of the magmatism (petroliferous
charge) was calculated, mapped, and interrelated with other processes (trapping, migration, and accu-
mulation) and essential elements (reservoirs and seal) to unravel the temporal and spatial relationships of
this atypical Permian petroleum system.
377
378 Araújo et al.
COLOMBIA
0º 0º
ECUADOR
CA-3
R-1
8º BRAZIL
PE
8º
RU
7250Km PT-1
CS-1
U
AY
24º 24º
PACIFIC
PR
RS-1
INA
PH-1
ENT
OCEAN
32º
N
ARG
32º URUGUAY A
OCE
C
NTI
UV-1
Outcrops of LA
Irati Fm. AT PU-1 CN-1
40º
40º
AL-1
MC-3
0 500Km RCH-1
48º
48º
SC
90º 78º 54º 42º 30º
TG-1
SE-1
7000Km
HV-1
MB-1
PI-1
0 70Km
400Km 650Km
Ponta Grossa
(A)
Rio Branco
do Sul
(B)
Palmeiras Curitiba
Campo Largo
Irati
Mandirituba
São Mateus area
AR-1
São Mateus do Sul
LA-1 LI-1
Três Barras area Campo do Tenente
DO-4 CB-3 TB-1 PP-1 AA-1
Rio Negro river PA-1
AT-1
Rio Negro MA-1 AB-1
er
Três Barras
riv
Mafra RI-1
AP-1 SA-1 PN-1 GU-3
u
aç
Canoinhas SJ-1
AN-1
u
Ig
1-CN-1-SC API-1
CS-2
LS-1
Outcrops of 0 30Km
Irati Fm. PH-1
Papanduva M-1A
UV-1
GO-1 PU-1
AL-1 MC-3
TP-1 TG-1
SE-1 HV-1
PI-1 MB-1
MA-1 LA-1
RD-1
ES-2 SJQ-1
LV-1
AO-1 MC-1
IT-1
AL-1
tions, the pattern of variation in these parameters was 20% in the transition zones. Similar results have been
determined in relation to different thicknesses of diabase observed by Simoneit et al. (1978, 1981), Dennis et al.
sills and the average depletion of the total organic carbon (1982), Peters et al. (1983), and Saxby and Stephenson
(TOC) content was calculated. (1987).
The behavior of geochemical parameters along the In the thermally altered zone, saturated to aromatic
eastern border was extrapolated for other areas of the hydrocarbon ratios increase and the isoprenoid/n-alkane
basin, where 48 wells were studied (Figure 1). On this ratios decrease. Furthermore, a shift toward the n-alkane
basis, it was possible to relate the results of the control distribution (higher concentration of n-alkanes with low
area with the rest of the basin, where the Assistência molecular weight) was observed, similar to the results of
Member of the Irati Formation extends across an area of Clayton and Bostick (1986).
approximately 700,000 km2. Biomarker maturity parameters derived from C32
Once we defined the area of calibration of the geo- hopane (22S/22S + 22R), C30 hopane and moretane
chemical and optical parameters and determined the per- (αβ/αβ + βα), Tm/Ts (trisnorhopane/trisnorneohopane),
cent average depletion of TOC caused by igneous intru- C29 ααα sterane (20S/20S + 20R), and C20 sterane
sive thermal effects, it was possible to reconstruct the (αββ/αββ + ααα) measured in samples with vitrinite
original TOC based on the medium residual TOC of the reflectance values lower than 1% Ro indicated progres-
48 selected wells and to estimate the original generation sive transformation toward isomerization equilibrium.
potential (S1 and S2) of the sections affected by the heat of However, when vitrinite reflectance was greater than 1%
the igneous intrusions. After calculating the original gen- Ro, a reversal trend of molecular maturity indices was
eration potential of the sections altered by the diabase observed, caused by low-maturity biomarker signatures
sills, it was possible to measure the residual generation that occur in samples exposed to rapid heating rates or to
potential in the laboratory (S1r and S2r) to determine the high absolute temperatures (as monitored by Raymond
mass of petroleum produced by the heat of igneous intru- and Murchison, 1992).
sion, the migrated mass of hydrocarbons, and the volume The most frequent activation energies of the Irati kero-
of hydrocarbons migrated per area. This petroleum gen range between 43 and 48 kcal/mol, with the preex-
charge was then related to the elements and processes ponential factor between 1.23 × 1010 and 2.13 × 1012,
that constitute the atypical Permian petroleum system of showing a transformation rate greater than 80% in the
the Paraná Basin. thermally altered zone. In some cases, a drastic decay in
all activation energy levels was observed.
The basic premise for geochemical calibration of the
data was based on the observation of a bulk depletion in
ASSESSMENT OF HYDROCARBON the organic carbon mass of the kerogen in the thermally
GENERATION CAUSED BY THE HEAT affected zone. This depletion is well defined by an almost
OF IGNEOUS INTRUSIONS complete transformation of the source rock generation
potential (S2). The mass of organic carbon consumed in
The thermal effect of igneous intrusions was observed the kerogen thermal degradation is proportional to the
in data derived from Rock-Eval, Fisher pyrolysis, optical volume of hydrocarbons generated by the system in liq-
parameters (Spore Color Index, or SCI), and vitrinite uid and/or gaseous form.
reflectance (% Ro) taken above and below intrusions in a The first step in reconstituting the original generation
thickness equivalent to that of the intrusive body, with an potential (S1 and S2) of the thermally affected zone was to
error of ±10%, also observed by Sad and Saraiva (1982). determine the distribution and thicknesses of the diabase
The transition zone between altered and unaltered source sills (Figure 2) and to correlate these with the depletion
rocks is abrupt when monitored by Fisher pyrolysis data percentages of the organic carbon caused by igneous heat
and transitional when based on SCI and % Ro data. input. For diabase sills with thicknesses equal to or
The original hydrogen index (HI) varies from 319 to greater than the thickness of the source rocks, a maximum
924 mg HC/g TOC, but the residual values of HI in the decrease of 30% was adopted for the TOC value, which
thermally affected zone are about 10% of the original corresponds to the average value observed in wells mon-
value and decrease to almost zero where the thickness of itored along the border of the basin.
the igneous intrusion is greater than the source rock thick- To make the TOC reconstitution possible and system-
ness. Such results were also observed by Cerqueira and atic and to decrease the error margins, the method con-
Santos Neto (1990) and Ferreira et al. (1994). Maximum siders that, despite facies variations, the system will be
TOC depletion (50%) was observed near the contacts depleted in organic carbon mass by an average of 30%.
with the igneous intrusions; TOC values gradually This value of 30%, based on the calibration of optical and
increase away from the intrusive contacts. Average TOC geochemical parameters (vitrinite reflectance, SCI, and
depletion in the transition zone was 30%. Similar results Rock-Eval pyrolysis data), supplied the correction factors
have been observed by Clayton and Bostick (1986). that were applied to the residual TOC for each interval
Nuclear magnetic resonance (NMR) indicates that the variation in the extent of the thermally altered zone,
aromatization of the kerogen decreases away from the whose thickness is equivalent to the igneous intrusive
intrusive contacts, varying from 100% at the contacts to thickness.
380 Araújo et al.
8200000
AG1
40
80
8000000
RA1
20
20
RP1
CG1 7750000
OL1
IL1
AR-1
LA1 LI1
PE1
SD1
40 60
DO3 AA1
CB1 TI1 TB1
PP1
MV1 PA1
PG1 AS1
100 AT1
MA1 AB1 7500000
60
40 40 AV1
AM1
RI1 SA1 GU3
CP1
AP1 PN1
CM1
0
22
CA3
RO1
RCA1
RV1
PT1
60
7250000
CS2
GP1
LS1
20
80
RS1
PH1 M1A
0 -20
20-40 RC1
40-60 UV1
PU1 CN1
0
60-80
10
BOUNDARIES
180 RD1 MA1 PA1
AO1
an
IT1
ce
20 TO1
cO
6750000
AL1
la nti
At
6500000
A
-100000 150000 400000 650000 900000
Boundaries of the Paran basin in Brazil 0 50 100 150 200 250km
35
30
Percentage of occurrence
25
20
(%)
15
10
0
0 0-5 5-0 10-20 20-30 30-40 40-50 50-60 60-70 70-80 80-90 90-100 100-110 110-120 120-130 130-140 140-150 150-160 >160
Figure 2—(A) Distribution of diabase isoliths and (B) isopach percentages in the Irati Formation.
Chapter 26—Atypical Permian Petroleum System of Paraná Basin, Brazil 381
After correction of the residual TOC in the original ual potential (S2r) (Figure 9) were used to quantify the
TOC, one section of the Assistência Member of the Irati volume of hydrocarbons removed from the source rocks
Formation was selected that had good correlation in response to the atypical generation caused by the heat
between the TOC and the generation potential for the of igneous intrusive bodies.
evaluation of the original generation potential (S1 and S2)
of the thermally affected section. In the selected section,
the relationship between the TOC content and the gener-
ation potential of the insoluble fraction (S2) presented a MASS BALANCE CALCULATIONS
degree of statistical correlation of 94%, defined by the The systematic calculations applied by Espitalié et al.
minimum square method. The relationship between the (1987) in Paris basin were adapted to Paraná basin.
TOC and the soluble fraction (S1) gave a correlation of Pyrolysis parameters were used to calculate the amount
87%. The generation potential and the soluble fraction of produced hydrocarbons and to monitor its distribution
calculated by the correlation equations of the standard in the Paleozoic charge system. The applied method can
well is given by be summarized in three stages:
S2c = 7.432 × TOC – 5.195 (R2 = 0.94) 1. Calculation of the mass of produced hydrocarbons
and (∆Sp), obtained by the difference between the orig-
inal genetic potential (S1o + S2o, which was calcu-
S1c = 0.4965 × TOC + 0.1275 (R2 = 0.87).
lated through reconstituted TOC using the correla-
tion equations between TOC versus S1 and TOC
The area where the optical and geochemical parame- versus S2 in the standard well) and the residual
ters were thermally affected by igneous intrusions was generation potential (S2r, which was measured in
designated as the thermal halo. To determine the the laboratory from the intervals altered by the
Assistência Member thermal halo, the direct relationship heat of intrusion) (Figure 10):
was used, and the extent of the thermally altered zone
was considered to be equivalent to the thickness of the ∆Sp = (S1o + S2o) – S2r (kg HC/ton rock).
intrusive.
The results plotted on a map provided the identifica- 2. Calculation of the mass of hydrocarbons produced
tion of two different areas (Figure 3). The first, corre- per unit of area (∆Sp/1 km2), in a section of 1 km2,
sponding to two-thirds of the basin area, includes with thickness (h) equivalent to the thickness of
Assistência Member sections with diabase sills having the Assistência Member (for each analyzed section,
thicknesses equivalent to or greater than the source with TOC > 1% and organic matter of type I + II)
rocks (thermal halo ≥100%). This area roughly coincides and constituted by rocks with an average density
with the area where the diabase intruded into the Irati of 2.5 ton/m3:
Formation is more than 20 m thick (Figure 3). The sec-
ond area, equivalent to the remaining third of the basin (∆Sp/km2) = ∆Sp × h × 2.5 × 103 (ton HC/km2).
area, corresponds to the area where the Irati section has
a thermal halo varying from 0% (absence of intrusions) 3. Calculation of the mass of hydrocarbons migrated
to 100%. per unit of area (Sm/km2), determined by the dif-
Only wells that revealed a section with a sufficient ference between the residual hydrocarbons (S1r)
number of reliable geochemical analyses (more than 10) and the generated hydrocarbons per unit of area
in the Irati Formation were used for the reconstitution of (∆Sp/km2):
TOC and original S1 and S2. Because of this restriction,
only 48 wells were selected. The scarcity of geochemical (∆Sm/km2 = S1r – (∆Sp/km2) (ton HC/km2).
data in the northern part of the basin made geochemical
reconstitution unfeasible in this area. To make use of the standard volumetric unit in the
After characterizing the percent thermal halo for each petroleum industry (m3), a conversion factor of 1.66 was
investigated section of the Irati Formation, relative cor- introduced that represents the average density of hydro-
rections were applied to calculate the depletion of TOC carbons (600 kg/m3) generated under a pressure of 20
from the sections where the residual mean TOC was MPa and a temperature higher than 200°C (Figure 11):
determined (Figure 4). After reconstitution of the resid-
ual TOC in each section affected by diabase sills (Figure
Volume of hydrocarbons (m3/km2) =
5), the original TOC was applied to the correlation equa-
tions (TOC versus S1 and TOC versus S2) obtained for 2.5 × ∆Sm × h × 103 × 1.66 (m3 HC/km2).
the original section. This was done to calculate the mass
of original free hydrocarbons (S1o) (Figure 6) and the According to Espitalié et al. (1987), the distribution of
original generation potential (S2o) (Figure 7). The recon- the hydrocarbon mass along the perimeter of the source
stituted genetic potentials compared with the residual rocks defines the compartmentalization of three different
mass of free hydrocarbons (S1r) (Figure 8) and the resid- areas: (1) the depleted area, where ∆Sm/km2 is negative,
8200000
AG1
TQ1
JA1
8000000
RA1
7750000
CG1 RP1 TL1
OL1
AR1
LA1 LI1
SD1
CB1 TI1 TB1 PP1 AA1 PA1
DO3 PG1
MV1
AT1 7500000
AB1
MA1
AV1
RI1 J1 SA1 GU3
AM1 CP1
AP1 JT1 PN1
SJ1
AN1
OT1
CM1 O1
API1
CA3
RP1
PT1 R1
7250000
LS1 GP1
M1A
RS1 PH1
UV1
PU1
AL1
GO1 MC3 ON1
TV1
RCH1 TP1
SE1 CA1
HV1 TG1 7000000
PI1 MB1
MR1
MA1 LA1
RD1 LV1 ES1
n
SJQ1
ea
RD1
Oc
MC1
AO1
tic
IT1
TQ1
lan
6750000
At
AL1
LEGEND
Thermal halo < 100%
Figure 3—Thermal halo map showing the effect of igneous intrusions on the Irati Formation.
Chapter 26—Atypical Permian Petroleum System of Paraná Basin, Brazil 383
8200000
AG1
TQ1
JA1
8000000
RA1
7750000
RP1 TL1
CG1
OL1
AR1
LA1 LI1
SD1 2
CB1 TI1 TB1 PP1 AA1 PA1
DO3 PG1
MV1
AT1 7500000
MA1 AB1
AV1
J1 SA1 GU3
RI1 PN1 CP1
AM1
AP1 JT1
SJ1
AN1
OT1
CM1 O1
API1 2
CA3
RP1
PT1 R1
7250000
LS1 GP1
M1A
RS1 2 PH1
UV1
PU1
AL1
ON1
GO1 MC3
3 TV1
RCH1 TP1
SE1 CA1
HV1 TG1 7000000
PI1 MB1
MR1 5
MA1 LA1
4
SJQ1
2
Oc
AO1
2-3% IT1
TQ1
lan
3-4% 6750000
At
1
4-5%
2
AL1
>5%
3
Figure 4—Average residual organic carbon distribution in the Assistência Member of the Irati Formation,
Paraná Basin (CI = 1).
384 Araújo et al.
8200000
AG1
TQ1
JA1
8000000
RA1
7750000
CG1 RP1 TL1
OL1
AR1
1 LA1 LI1
SD1 2
CB1 TI1 TB1 PP1 AA1 PA1
DO3 PG1
MV1
AT1 7500000
4
AB1
MA1
AV1
3 J1 GU3
RI1 SA1
AM1 CP1
AP1 JT1 PN1
SJ1
AN1
OT1
CM1 O1
API1
CA3
RP1
PT1 R1
7250000
LS1 GP1
M1A
1 RS1 PH1
UV1
PU1
AL1 3
GO1 MC3 ON1
TV1
4 RCH1
TP1
SE1 CA1
HV1 TG1
7000000
PI1 MB1
MR1
MA1 LA1
RD1 LV1 ES1 5
n
5 SJQ1
ea
0-1% 4
Oc
RD1 MC1
3
1-2% AO1
tic
IT1
2-3% TQ1
lan
6750000
3-4% 2
At
3
4
4-5% AL1 5
>5%
Figure 5—Reconstituted average organic carbon distribution in the Assistência Member of the Irati Formation,
Paraná Basin (CI = 1).
Chapter 26—Atypical Permian Petroleum System of Paraná Basin, Brazil 385
8200000
AG1
TQ1
JA1
8000000
RA1
7750000
RP1 TL1
CG1
OL1
AR1
LA1 LI1
SD1
CB1 TI1 TB1 PP1 AA1 PA1
DO3 PG1
MV1 2 AT1 7500000
MA1 AB1
AV1
RI1 J1 SA1 GU3
AM1 CP1
AP1 JT1 PN1
SJ1
AN1
OT1
CM1 O1
API1
2
CA3
RP1
PT1 R1
7250000
LS1 GP1
M1A
RS1 PH1
UV1
PU1
AL1
GO1 MC3 ON1
TV1
RCH1 TP1
3
SE1 CA1
HV1 TG1
7000000
PI1 MB1
MR1
MA1 LA1
RD1 LV1 ES1
eeaan
n
2 SJQ1
0-1
OOcc
RD1 MC1
1-2 1 AO1
ttiicc
IT1
2-3 TQ1
laann
6750000
>3
AAttl
2
AL1
00 50
50 100
100 150
150 200
200 250km
250km
6500000
Figure 6—Original mass of free hydrocarbons in the Assistência Member of the Irati Formation, reconstituted through an
equation of correlation (S1o = kg HC/ ton rock; CI = 1).
386 Araújo et al.
8200000
AG1
TQ1
JA1
8000000
RA1
7750000
RP1 TL1
CG1
OL1
AR1
LA1 LI1
SD1 10
CB1 TI1 TB1 PP1 AA1 PA1
DO3 PG1
MV1
20 AT1 7500000
MA1 AB1
AV1
RI1 J1 SA1 GU3
AM1 CP1
AP1 JT1 PN1
SJ1
AN1
10 OT1
CM1 O1
20
API1
CA3
RP1
PT1 R1
7250000
LS1 GP1
M1A
RS1 PH1
UV1
PU1
AL1
GO1 MC3 ON1
TV1
RCH1 TP1
SE1 CA1
HV1 TG1 7000000
PI1 MB1
MR1
MA1 LA1
RD1 LV1 ES1
30 SJQ1
n
ea
10-20 AO1
tic
IT1
20-30 TQ1
lan
6750000
30-40
10
At
20
40-50 AL1 30
>50
Figure 7—Average original source rock potential in the Assistência Member of the Irati Formation, reconstituted through an
equation of correlation (S2 = kg HC/ ton rock; CI = 10).
Chapter 26—Atypical Permian Petroleum System of Paraná Basin, Brazil 387
AG1
TQ1
JA1
RA1
RP1 TL1
CG1
OL1
AR1
LA1 LI1
SD1
CB1 TI1 TB1 PP1 AA1 PA1
DO3 PG1
MV1
AT1
MA1 AB1
AV1
RI1 J1 SA1 GU3
AM1 CP1
AP1 JT1 PN1
SJ1
AN1
OT1
CM1 O1
API1
CA3
RP1
PT1 R1
LS1 GP1
M1A
RS1 PH1
UV1
PU1
AL1
GO1 MC3 ON1
TV1 3
RCH1 TP1
SE1 CA1
HV1 TG1
PI1 MB1
MR1
MA1 LA1
RD1 LV1 ES1
SJQ1
0-1
RD1 MC1
n
1-2
ea
AO1
Oc
IT1
2-3 TQ1
tic
3-4 1
lan
4-5 2
AL1
At
>5
Figure 8—Residual mass of free hydrocarbons in the Assistência Member of the Irati Formation (S1r = kg HC/ ton
rock; CI = 1).
388 Araújo et al.
8200000
AG1
JA1
TO1
8000000
RA1
RP1
7750000
CG1
OL1
TL1
AR1
LA1 LI1
PE1
SD1
0
DO3 CB1 TB1 AA1
TIL
PP1
MV1 PA1
25
PG1AS
AT1
15
MA1
7500000
AB1
5
20
AM1 AV1
J1
SA1
RI1 5 GU3
PN1 CP1
AP1
JT1
SJ1 10 QT1
AN1
API1 15
CM1
MO1
CA3
10 RO1
RV1 R1
RCA1
PT1
RP1 CS2
7250000
15
GP1
LS1
RS1
PH1
M1A
RC1
UV1
CN1
PU1
GO1
AL1
MC3
RCH1
TP1 TV1
CA1
TG1
SE1
HV1
MB1
7000000
PI1
BN1
0-5 MA1
5 - 10 LA1 5
PA1
MR1 0 5
10 - 15 RD1
20
15 - 20
25
20 - 25 LV1 ES1 MC1
30
25 - 30 SJ01
30 - 35 RI1 5 15 10
35 - 40
A01
an
IT1
ce
TO1
cO
10
6750000
15
nti
AL1
20
la
25
At
6500000
Figure 9—Average residual source rock potential in the Assistência Member of the Irati Formation (S2r = kg HC/ ton rock; CI
= 5).
Chapter 26—Atypical Permian Petroleum System of Paraná Basin, Brazil 389
8200000
AG1
JA1
TO1
8000000
RA1
RP1
CG1
7750000
OL1
TL1
AR1
LA1
LI1
PE1
SD1
DO3 AA1
CB1 TB1
TI1 PP1
MV1 PA1
PG1
AT1 AS1
MA1
7500000
AB1
20
10
AM1 AV1
10
J1 SA1
RI1
PN1 CP1 GU3
AP1
JT1
SJ1 QT1
AN1
10
API1
CM1 01
20
MO1
CA3
RO1
RV1 R1
RCA1
RP1 PT1
CS2
7250000
GP1
LS1
RS1
PH1
M1A
RC1
PU1 CN1
GO1
AL1
MC3
RCH1
TP1
CA1
20
TV1
TG1
SE1
HV1
7000000
MB1
PI1
30
BN1
MR1 MA1
LA1
0 - 10
10
RD1
PA1
10 - 20
20 - 30 LV1 ES1
SJO1
30 - 40 RI1 MC1
30
AO1 20
n
IT1
ea
TO1
Oc
10
6750000
tic
AL1
lan
At
6500000
Figure 10—Distribution of the hydrocarbon mass produced by the thermal effect of igneous intrusives on the Assistência
Member of the Irati Formation (∆Sp = S2o + S1o – S2r kg HC/ ton rock; CI = 10).
390 Araújo et al.
8200000
AG1
JA1
TO1
RA1
RP1 7750000
CG1
OL1
TL1
AR1
LA1
LI1
PE1
SD1
100
AV1
0
AM1
J1
RI1 SA1
CP1 GU3
500
AP1 PN1
JT1
QT1
SJ1
AN1
500
API1
CM1 O1
MO1
CA3
RO1
RV1 R1
RCA1
GP1
LS1
0 - 500
RS1
500 - 1000 PH1 M1A
1000 - 1500
RC1
UV1
1500 - 2000
PU1 CN1
2000 - 2500 GO1
AL1
2500 - 3000 MC3 00
RCH1 20
3000 - 3500 TP1
00
TV1
10 CA1
TG1
25
SE1
00 3000HV1
MB1
PI1
MR1 BN1
MA1
15 LA1
00
RD1 PA1
-30
00
2000
LV1 ES1
SJO1
RI1
MC1
0
50 0 2500 00
100 AO1 10
IT1
1500 0
50
TO1
AL1
6500000
Figure 11—Areal distribution of the volume of hydrocarbons migrated due to the thermal effect of igneous intrusives on the
Assistência Member of the Irati Formation (V m3/km2 = 2.5 × 1.66 × ∆Sm × h m3 HC/km2; CI = 500 × 103).
Chapter 26—Atypical Permian Petroleum System of Paraná Basin, Brazil 391
indicating lateral or vertical migration outside of the hydrocarbons from the source rocks toward the carrier
source rocks; (2) the accumulation area, where ∆Sm/km2 beds). This demonstrates that the maximum expulsion
is positive, indicating lateral migration to the bordering occurs when the pressure reaches values between 300 and
areas of the generating system; and (3) the nonmigration 500 kgf/cm2. As a consequence of internal pressurization
area, where ∆Sm/km2 is equal to zero, indicating that the of the source rocks, geopressured compartments origi-
generated mass remained in the source rocks. nate in connection with the development of diagenetic
The charge calculated by the equations, that is, the vol- seals above the generation kitchens. In most of the stud-
ume of available hydrocarbons to be trapped (Sluijk and ies published by Ortoleva (1994), the geopressured com-
Nederlof, 1984) varied from zero to 3500 × 103 m3/km2 partments were related to the phase of kerogen thermal
(Figure 11). The volume of migrated hydrocarbons (m3) cracking.
per area (km2) was delineated in the geochemical map by Some evidence of geopressurization in the Paraná
curves of isocharge of 500 × 103, with maximum isopletes Basin, attributed both to kerogen thermal cracking and to
of 3500 × 103. cracking of oil to gas, were preserved, in spite of further
The calculated charge is more expressive south of the intense tectonic compartmentalization. Geopressured
Ponta Grossa arch (fault zone of the Rio Alonzo), it being zones were detected in two wells drilled in the basinal
the maximum isocharge distributed in Santa Catarina depocenter. In one well, the existence of a geopressured
and Rio Grande do Sul states (Figure 11). Here, the source compartment was confirmed (1.22 kgf/cm2/m, 1.74
rock is richer and the thickness of the intrusives is greater psi/m) owing to a diagenetic seal immediately above the
than the source rock thickness. In São Paulo State, the Irati Formation. The pressure shift between the compart-
maximum isocharge of 1500 × 103 was mapped in a ments normally and abnormally pressurized above and
restricted area. This decrease in the charge was due to a below the seal is 85 kgf/cm2 (Figure 12). The geopressur-
decrease in the organic content and to the small thickness ized compartment with a pressure gradient of 1.22
of the diabase sills, mainly in the eastern part of the state. kgf/cm2/m was characterized from the Irati Formation
Basically, the configuration and location of the Irati (Kazanian) to the Ponta Grossa Formation (Devonian).
Formation kitchen reflects the spatial distribution of the The diagenetic seal above the Permian source rocks
diabase sills that affected the organic content contained in provides evidence that the thermobaric reactions in the
the bituminous shale and marls. In comparative terms, source rocks supplied dissolved silicate and carbonate for
the migrated maximum amount of the Irati section (3500 cementation, similar to what has been observed in other
× 103 m3/km2) converted to the source potential index basins (Hunt, 1990; Al-Shaieb et al., 1994; Bradley and
(SPI = 2.1 ton HC/m2), established by Demaison and Powley, 1994; Shepherd et al., 1994; Surdan et al., 1994).
Huizinga (1991) to classify the available charge in vertical The seal was preserved for 127 million years, probably
drainage, would be considered abnormally low to supply because of the basin center location of the area, where the
the trapped subsystems effectively. effects of uplift common along the borders were attenuat-
The discrimination of the type of hydrocarbons gener- ed. The occurrence of Rio Bonito Formation sandstones in
ated (liquid or gaseous) by intrusive heating constitutes the depocenter with porosities of 20% at 4000 m of depth
an unsolved problem in the evaluation of the petroleum would also indicate that acidic fluids coming from the
charge. The maturity level based on geochemical and Devonian source rocks dissolved cement and created sec-
optical parameters suggests that the measured thermal ondary porosity.
stress is compatible with the liquid hydrocarbons win- The great thickness of the geopressurized compart-
dow. However, there is no parameter that can be used to ment (~2500 m from Permian to Devonian source rocks)
infer the speed of the intrusive emplacement or the speed is possibly due to the pressure equalization that occurred
of primary migration in order to assess the relative between the two generating systems. The Devonian sys-
amount of liquid hydrocarbons cracking to gaseous tem pressurization was caused by the thermal cracking of
hydrocarbons in the source rocks. The drastic hydrogen oil to gas during magmatism. In a sealed system, the
loss (>90%) in a thermal maturity level equivalent of 0.7 cracking of 1% of oil can be enough to increase the pore
and 2% Ro can be attributed to the direct generation of pressure from the hydrostatic to the lithostatic limit
high amounts of gaseous hydrocarbons. The quantitative (Gaarenstroom et al., 1993).
approach used in this paper considered the petroleum The relationship between the residual mass of free
charge generated only as liquid hydrocarbons. hydrocarbons (S1r) present in the shale and marls of the
Irati Formation (Figure 8) and the mass of produced
hydrocarbons (S2o + S1o – S2r) by the thermal igneous
effect (Figure 10) indicates that the efficiency of expulsion
GEOPRESSURING AND PRIMARY of hydrocarbons from the source rocks was almost 100%.
MIGRATION This high efficiency, greater than that verified by Cooles et
al. (1986), Mackenzie et al. (1987), Espitalié et al. (1987),
Experiments conducted by Lafargue et al. (1994) on the Larter (1988), and Pepper (1991) (75–90%), can be attrib-
lower Toarcian shales of Paris Basin have highlighted the uted to the igneous promotion of secondary cracking (oil
importance of geopressurization of the source rocks as an to gas) of the remaining liquid hydrocarbons that are nor-
essential factor in primary migration (expulsion of the mally retained in the organic matter or in the rock miner-
392 Araújo et al.
1000
Normal gradient of the area: 1.47 psi/m
compartment
2000 with normal
pressure
DEPTH (M)
6000
0 100 200 300 400 500 600
PRESSURE (KGF/CM2)
Figure 12—Geopressurized compartment observed at well 1-API-1-PR (central Paraná Basin) with a diagenetic seal above
the Permian source rock (Irati Formation). Pressure data obtained from cable tests (RFT).
al matrix. However, gaseous hydrocarbons cannot be cor- observed near the contact between the intrusion and
rectly measured during drilling. Sokolov et al. (1971) con- source rock are interpreted to reflect the overlap of frac-
cluded that the sampling of source rocks with preserva- turing caused both by hydrocarbon generation and inser-
tion of original temperature and pressure conditions con- tion of the sill.
tain 98% more gas than the samples collected during con- To evaluate the primary migration efficiency, relative
ventional drilling. quantification of residual bitumen was done on thin sec-
Qualitative analysis of the hydrocarbon occurrences tions from two Irati sections intruded by sills of 28.4 m
sampled during drilling of the Irati Formation indicates and 2 m thickness, respectively. In the well with the thick-
that, in the area with a thermal halo equal to or greater er sill, rare occurrences of residual bitumen were
than the source rock thickness (thickness of intrusions ≥ observed, while in the well with the thinner sill, larger
source rocks), no occurrence of liquid hydrocarbons was amounts were observed. The reverse relationship
reported. This corroborates the hypothesis of complete between relative abundance of residual bitumen in
cracking of the oil remaining in the source rocks to source rocks and thickness of the sills indicates that the
gaseous hydrocarbons. In wells more recently drilled in efficiency of the primary migration is directly related to
the basin, which used special muds and more accurate heat dissipation from the sills. Thinner sills would allow
detection equipment, gas was detected during penetra- faster reestablishment of the thermal balance, reflected in
tion of the Irati Formation. These observations reinforce decreased efficiency of the primary migration.
the hypothesis that the expulsion efficiency from the Irati Analysis of the homogeneization temperatures (Th)
Formation has been similar to the level observed in other measured in aqueous and oil fluid inclusions (Fuzikawa,
basins (75–90%). 1994) within calcite and quartz veins cutting two Irati sec-
A petrographic investigation of microfractures in thin tions intruded by diabase sills 27.2 m and 20.6 m thick,
sections from a well affected by a 27.2-m-thick diabase sill respectively, allowed the following inferences:
revealed that 90% of the microfractures were distributed
parallel to bedding and were partially filled with calcite. 1. In the well affected by the thinner sill, water inclu-
Quantitative analysis of the relative abundance of sions with Th ranging from 40° to 240°C indicate
microfractures indicates a high concentration of horizon- that the filling of the fractures began during the
tal fractures at about 9 m from the contact with the generation pulses and continued during uplift of
igneous body and another increase at about 18.5 m from the basin borders started in the Cretaceous. The
the basal contact. This reinforces the hypothesis that the temperatures of the oil inclusions, varying from
microfractures are more likely related to thermobaric 60° to 260°C, indicate that during primary migra-
causes due to kerogen cracking than to mechanical shear tion, hydrocarbon pulses were displaced under
caused by intrusion of the sill. The open microfractures different temperatures. This results from the differ-
Chapter 26—Atypical Permian Petroleum System of Paraná Basin, Brazil 393
ence between the maximum level of heat in the form a mosaic whose limits correspond to zones of
source rocks and the cooling that occurred during weakness and pressure relief, where the primary migra-
migration. The higher homogeneization tempera- tion converged away from intrusion sites during mag-
tures were recorded in samples closer to the sill, matism.
both for aqueous and oil inclusions, which is con- Most of these megafeatures were reactivated during
sistent with the temperature gradient expected the Cretaceous magmatic event due to forces mainly
from contact with the intrusion. along the northwest faults zones, where an intricate net-
2. In the well intruded by the thicker sill, the aqueous work of dikes was intruded. This occurred in connection
inclusions showed homogeneization temperatures with the transtensional movements caused by the initial
ranging from 40° to 150°C. The sample closest to stages of Gondwana’s breakup and the vertical uplift of
the igneous body showed mainly aqueous inclu- the sedimentary section caused by the insertion of hori-
sions and some oil inclusions with homogeneiza- zontal and subhorizontal sills that form about 12% of the
tion temperatures from 50° to 70°C. A possible total column of the basin (O. A. Zanotto, personal com-
explanation for this is the intense shearing and munication, 1998). According to Ar/Ar dating (Turner et
thermal expansion that occur during intrusion. al., 1994), the Serra Geral magmatism occurred between
During the cooling of the sill, the pressure relief 137 and 127 Ma. Milani (1997) inferred from Ar/Ar ages
caused by the open fractures caused a reversion in from of a well drilled in the basin depocenter that both
the direction of the flow, carrying hydrocarbons intrusive and extrusive magmatism occurred simultane-
toward the contact zone which had maturity geo- ously between 138 and 128 Ma.
chemical parameters (biomarkers) incompatible The large vertical movements of blocks, creating a
with the optical parameters (% Ro and SCI), as ver- tectonic megacompartmentalization, generated zones of
ified by Trigüis (1986) and Mendonça Filho (1994). pressure relief where the flows converged. The frame-
The immature parameters of the biomarkers indi- work of this megacompartmentalization was broken
cating thermal maturity (Ts/Ts + Tm), tricyclics/ into smaller blocks, creating multiple areas of pressure
pentacyclics, C32 homohopanes (22S/22S + 22R), relief of a few square kilometers. Investigation of four
C29 steranes (20S/20S + 20R), and C29 steranes areas in Guiabá Paulista (SP), Taquara Verde (SC), Três
(αββ/αββ + ααα) were determined along the sec- Pinheiros (SC), and Matos Costa (SC) which had avail-
tions affected by the sills in both wells. They regis- able seismic control and closely spaced wells, confirmed
ter a reverse primary migration of less thermally the local compartmentalization at the level of the
evolved hydrocarbons toward the contact, as well Permian source rocks, as shown in Figure 14, which
as migration of thermally evolved hydrocarbons reveals bordering faults at about every 2 km. The sills
away from the contact. mapped into the source rocks vary in thickness and
3. The trapping of innumerous oil inclusions with stratigraphic level, indicating that the vertical buoyancy
homogeneization temperatures from 50° to 130°C of the blocks activated the bordering faults of the com-
in the Palermo Formation, 15.5 m below the base partments, creating zones of pressure relief to which pri-
of the Irati source rocks, indicates that the vertical mary migration converged.
propagation of the fractures extended beyond the The activation of each compartment probably
limits of oil generation. This reinforces the theory occurred in a sequential way, considering the time and
that petroleum can migrate vertically both upward spatial differentiation in the insertion of the igneous intru-
or downward from the source rocks when the dif- sions. The resulting sequential mechanism of local pres-
ferential pressure is great enough. The capture of sure relief transformed the generation kitchen in a mosa-
oil inclusions under a high-temperature spectrum, ic composed of several separate and unsynchronized
in the section underlying the Irati Formation indi- kitchens. From this, it can be concluded that primary
cates that, during the multiple pulses of migration, migration occurred mostly laterally along short distances
the fractures were kept opened by the geopressur- (1–10 km) to the bordering faults of the compartments,
ization of the generation system. which were the conduits for vertical migration to the
sandstone reservoirs.
The efficiency of the primary migration process in the Because of the contemporaneity of the generation and
Irati Formation can be attributed to the existence of a primary migration, the countless intrusions dated
network of predominantly horizontal fractures and between 127 and 138 Ma (Turner et al., 1994; Milani, 1997)
microfractures and to the extensive tectonic compart- caused multiple generation–migration pulses from differ-
mentalization of regional and local scale. This facilitated ent kitchens. These had a high expulsion efficiency due to
the contemporaneity between primary and secondary the creation of a network of horizontal microfractures that
migration through activation of the faults that bordered interconnected the geopressurized kitchens to the vertical
the compartments. The structural framework, reinter- pressure relief faults (Figure 15). Due to this sequence of
preted by Marques et al. (1993), matched with the geo- events, it can be imagined that great losses of petroleum
chemical map of the petroleum charge, shows that the charge occurred due to an excess of vertically focused
Permian kitchen is segmented by 20 structural megafea- flow and to fragmentation of the petroleum charge that
tures (rupture lines and faults zones) (Figure 13). These reached the rock reservoir in intermittent pulses.
394 Araújo et al.
8200000
AG1
JA1
TQ1
8000000
RA1
CG1
RP1 7750000
OL1
TL1
AR1
LA1 LI1
PE1
SD1
10
00
0
AV1
AM1
J1
SA1
RI1
CP1 GU
500
AP1 PN1
JT1
QT1
5J1
AN1
500
API1 CM1 01
MO1
CA3
RQ1
RV1 R1
RCA1
RP1
CS2 PT1 7250000
GP1
LS1
RS1 PH1
0 - 500
M1A
500 - 1000
1000 - 1500
UV1 RC1
1500 - 2000
PU1
2000 - 2500 GO1
CN1
MC3
3000 - 3500
20
RCH1
TP1
TV1 00
VOLUME OF 10SA1
MIGRATED HYDROCARBON TG1
SE1
PER AREA 25 30 HV1 7000000
00 00
(m3HC/km2) PI1
MB1
MR1 BN1
15 MA1 LA1
00 A1
RD1
30
2000
00
LV1 ES1
SJ01
RI1 MC1
FAULT ZONES LINEAMENTS
0
100
0 2500 06 01
00
FARTURA TRANSBRASILIANO
50 10 AO1
07 S. JERÔNIMO/CURIUVA 02 UBERLÂNDIA
IT1
1500 09 C. DE ABREU 04 GUAXUPÉ
10 LANCINHA/CUBATÃO 05 BOTUCATU
TO1
11 RIO PIQUIRI 13 RIO IGUAÇU 6750000
12 CAÇADOR 14 RIO CLARO
20 LEÃO 22 TAZAQUARA
23 MISSIONES/APUCARANA 30 JACUTINGA
24 CRUZEIRO DO OESTE
25 LOANDA/P.EPITACIO
26 ELDORADO/T. LAGOAS
27 AMAMBAI
28 MOGI GUAÇU/DOURADOS
6500000
CENTRAL PARANÁ AREA
Figure 13—Structural framework of the Paraná Basin and the Permian hydrocarbon kitchen resulting from the thermal
effect of diabase sills on the Irati Formation (CI = 500 × 103).
Chapter 26—Atypical Permian Petroleum System of Paraná Basin, Brazil 395
1-TV-3-SC
1-TV-2-SC
1,9k
m
2km m
2-TV-1-SC 2,1k
1-TV-4-SC
20
Fault mapped based on seism
Fault mapped ba
Diabase sill of 27m
30
Diabase sill of 30m
data
Diabase sill of 27m
sed
mic
40 Kitchen IV
on seismic data
Fault mapped based on seis
ic data
50
Kitchen II
Assistência Mb. Base
60
Figure 14—Spatial positioning of the diabase sills of the Irati Formation in the Taquara Verde area (SC) . The thickness
range and the insertion of different stratigraphic levels in restricted areas are indicative of the complex distribution of
igneous intrusives and the activation of compartmented kitchens of generation.
1 to 10km
LT
E FAU
HALO
UCTIV
THERMAL 1º BITUMINOUS SHALE
ZONE
COND
PREFERENTIAL
DIABASE SILL MARL FLUX DIRECTION
SILTSTONE
2º BITUMINOUS SHALE
TOC 30%
S2 ZERO
MARL
MICROFRACTURES
Figure 15—Generation and primary migration model of the Permian petroleum system of Paraná Basin fractures and
microfractures resulting from thermal cracking of the Irati Formation kerogen due to the effects of magmatic intrusion. The
preferential pattern of horizontal fractures was observed in thin sections.
2. The second phase, concomitant to the genera- Botucatu and Pirambóia Formations (Jurassic–Triassic)
tion–migration process (127–138 Ma), occurred that overlay the Permian source rocks in the Paraná Basin
during the active rifting of Gondwana. It was relat- constitute carrier beds favorable to long-distance migra-
ed to transtensional forces overlapped on the pre- tion. However, the fragmentation of the petroleum charge
vious structural framework imprinted in the supplied to secondary migration made displacements
Jurassic–Triassic section (Pirambóia and Botucatu along long distances unfeasible. Larger continuous phas-
Formations) creating complex structural inversions es of hydrocarbons were needed to compensate for losses
difficult to characterize in seismic mapping. during migration. In addition to this restriction, the exten-
sive tectonic compartmentalization tended to focus the
The contemporaneity of the generation–migration flow vertically.
process with the second phase of trap formation decreas- The only known hydrocarbon occurrences in the
es the potential for the Pirambóia and Botucatu Jurassic–Triassic sequence are found in an area of 7000
Formations to trap petroleum because of uncertainties km2 in the eastern part of the basin (São Paulo state),
about the synchronism between events. Based on this where 19 important shows of residual biodegraded
premise, it is reasonable to suppose that the greatest hydrocarbons in Pirambóia fluvial-eolian sandstones
chance of storage of petroleum charge occurred in the were mapped (Figure 16). The geochemical signature of
structures originated in the Cape–La Ventana event. these hydrocarbons indicates a Permian origin (Cerqueira
Lateral migration along long distances demands spe- and Santos Neto, 1986). In this area, the orogenic Cape–La
cial geologic conditions, such as the presence of reservoir Ventana event produced positive movements (Milani,
rocks of regional extent or migration routes through 1992) and an erosive episode that created a unconformity
regional unconformities. This has been verified in the surface on which rests the Pirambóia reservoir a distance
Alberta Basin of Canada, where migration of about 100 of 50–150 m from the Permian source rocks (Figure 16).
km has been suggested for the Athabasca tar sand (Deroo As the basal Teresina Formation is mainly shaly in this
et al., 1977). Due to their continental extent (839,000 km2) area, the first available carrier beds are the Pirambóia
(Araújo et al., 1995), fluvial-eolian sandstones of the sandstones. The petroleum charge of 1000 × 103 m3
TB-1 PP-1 AA-1 MA-1 RB-1 AB-1 AT-1 PG-1 AS-1
7km
80km 60km 97km 112km 22,4km 41km 22km
SEA LEVEL
0
ÓIA
MB .
RA Gp
U / PI ARÉ
CAT ITARdivided
U n
O TU INA
-500 sB RES
Fm . TE
Fm
-1000
asto
do R
Rio Inferred Fault
Fm
.
IRATI Fm p.
Fm.
SE RRA ALTA ÉG m.
-1500 o Fms.
R AF
o/Palerm RA OS
Rio Bonit ITA A GR
N T
PO
Fault mapped
m. through seismic data
nas F
-2000 Fur
G p.
Ivai Guaxupé fault
Rio
-2500
BASEMENT
LOCATION MAP
-3000 22o AA-1
TB-1 PP-1
AS-1
. AT-1
Fm MA-1 RB-1
s a Fm.
ros Furnas Mogi-Guaçu/Dourados Lineament PG-1
-3500 ta G
Pon AB-1
p.
IG
IVA
-4000 RIO Mogi-Guaçu/Dourados Lineament 24o
50o 48o
Unconformity
-4500
Guapiara Fault
Chapter 26—Atypical Permian Petroleum System of Paraná Basin, Brazil
Figure 16—Geologic section showing the Permian–Triassic unconformity (Cabo–La Ventana orogeny) which created the close
proximity of Permian source rocks with the Pirambóia Formation sandstones in the area of tar sand occurrences.
397
398 Araújo et al.
HC/km2 calculated for this area was most likely focused Two characteristics of this atypical Permian petroleum
directly toward this carrier bed instead of migrating lat- system support the hypothesis that lateral migrations
erally. The inferred residual volume of 60,000 × 103 m3 HC took place over a short distance: (1) the excessive number
(Hettich, 1981) represents approximately 1% of the pro- of vertical faults that focused flow vertically, and (2) frag-
duced volume, considering the area of 7000 km2. An mentation of the petroleum charge caused by multiple
amount of 0.25 m3 HC/m3 of sandstone, measured in the generation–migration pulses.
main occurrences of the tar sands, indicates that these
areas correspond to residual petroleum accumulation
biodegraded by meteoric gravitational flow during the
marginal uplift that began in the middle Cretaceous. TERTIARY MIGRATION
Calculation of a volumetric balance demands a deter-
mination of the residual hydrocarbon saturation after sec- The uplift of the borders of the Paraná Basin related to
ondary migration. In most carrier beds, the residual the South Atlantic rifting added to the propagation of
hydrocarbon saturation after secondary migration varies compressional stresses from the western Andean oroge-
from 1 to 10% (England et al., 1987). Macgregor and ny. This created a structural asymmetry that resulted in
Mackenzie (1987) determined the residual hydrocarbon the regional basculation of the eastern border toward the
saturation in the Sumatra Basin to be about 2%. Larter western, northern, and southern borders of the basin.
and Hortad (1992), studying the reservoirs of the East Apatite fission trace data indicate that the first uplift puls-
Shetland Basin, concluded that when the residual hydro- es of the eastern border began 110 Ma (Lelarge, 1993),
carbon saturation is about 3%, the balance between the reaching a climax between 100 and 90 Ma and causing
petroleum charge and the lost volume is positive and erosion of 3000 m of section. Zanotto (1993) also suggests
petroleum accumulations exist along the migration route. 3000 m of erosion using the vitrinite reflectance method.
When the residual hydrocarbon saturation is about 6%, The erosion intensified during the Cenozoic, causing the
the mass balance is negative and no hydrocarbon accu- erosional scarp to retreat more than 80 km from the orig-
mulations exist along the migration route. inal border, which provided the onset of the gravitational
Calculation of the volume of petroleum lost (Vp) along hydraulic head contemporaneous with the tertiary
the routes of migration is given by the equation Vp = φfVr, migration.
where φ is porosity, f is the residual hydrocarbon satura- Due to gravitational drive, the speed of water dis-
tion, and Vr is the rock volume crossed by the flow placement into the Botucatu and Pirambóia sandstones
(Mackenzie and Quigley, 1988). The difference between (with 17% average porosity) can reach values of about
the petroleum charge and the lost volume corresponds to 1000 km/0.22 m.y. (based on δ13C data from Silva, 1983).
the volume trapped. As there is no measurement of resid- This gravitational water movement is 4.5 times faster
ual hydrocarbon saturation in the Paraná Basin, the cut- than liquid hydrocarbon displacement via buoyancy
offs (3% and 6%) determined by Larter and Hortad (1992) (assuming a carrier bed with permeability of 1000 md;
were adopted to proceed calculation of the volumetric England et al., 1991) and resulted in the onset of the
balance of the petroleum charge (m3 HC/km2) using the Paraná Basin hydrodynamic traps.
above equation. In the Catarinense Plateau area along a 150-km extent
Two areas with significant petroleum charge were of the Peixe River, dozens of oil occurrences have been
selected to calculate the volumetric balance: the detected in the basalt from seeps along the railroad and in
Catarinense Plateau and north-central Paraná Basin close countless shallow wells. Oils of up to 33° API have been
to the basin depocenter (Figure 13). In the Catarinense collected here and geochemically correlated to the
Plateau area (SC), where the average petroleum charge is extracts of Permian source rocks. In these shallow wells,
1500 × 103 m3 HC/km2 and φ =14% (Teresina and Rio oil occurs in the fractured and vesicular zones at the tops
Rasto Formations), it was calculated (assuming f = 6%) of the lava flows, usually near the topographic surface
that the hydrocarbon charge would be dispersed (lost) in (~100 m depth). The incipient degree of bacterial degra-
a thickness of 178 m. Assuming f = 3%, φ = 20%, and So = dation provides evidence of recent tertiary migration ver-
70% (oil saturation), the storage thickness would be 6.4 m. tically from the Permian–Jurassic section. Intense renew-
The supply through the faults started from the Pirambóia al of meteoric water is active in the basaltic aquifer.
Formation base and migrated vertically 100 m up to the Long-distance lateral tertiary migration (<80 km) in
base of the Serra Geral Formation, a sealing package of the Paraná Basin is quite unlikely because of three factors:
volcanic rocks as thick as 2000 m.
For the north-central Paraná Basin, where the average 1. The existence of an intricate network of dikes, sills,
petroleum charge is 500 × 103 m3 HC/km2, it was calcu- and fault zones indicates short-distance migrations
lated that this charge would be dispersed in a thickness of with a tendency for vertical migration.
59 m. For a vertical buoyancy of 150 m up to the base of 2. The small thickness of possible accumulations in
the volcanic seal, the volumetric balance would be nega- the Jurassic–Triassic sections calculated from volu-
tive due to the lost volume of 900 × 103 m3 HC/km2, metric balance (<10 m) would be retained in the
which is larger than the calculated petroleum charge of pores as residual fraction, only a few hundreds of
500 × 103 m3 HC/km2. meters from the trap.
Chapter 26—Atypical Permian Petroleum System of Paraná Basin, Brazil 399
Events Key
400 300 200 100 m.y. GEOLOGIC
I. Source rock
TIME 1. Assistencia Mmbr of Irati Fm.,
PALEOZOIC MESOZOIC CENOZOIC Kazanian age
PETROLEUM II. Reservoirs
SYSTEM 2. Palermo and Rio Bonito ss
D C P TR J K PE N 3. Teresina and Rio do Rasto ss
EVENTS
4. Pirambóia and Botucatu ss
1 I - Source rock III. Seals
5. Palermo and Rio Bonito sh
2 3 4 II - Reservoir 6. Teresina and Rio do Rasto sh
7. Serra Geral igneous rocks
5 6 7 III - Seals IV. Burial
8. Late Permian–Early Triassic
8 9 10 11 IV - Burial 9. Late Triassic–Jurassic
10. Early Cretaceous
12 13 V - Trap formation 11. Late Cretaceous
V. Trap formation
VI - Generation/migration 12. La Ventana orogeny
14
and accumulation 13. Rifting phase
VII - Preservation VI. Generation/migration and
accumulation
VIII - Critical moment 14. Igneous intrusives thermal
15 16 effect
VII. Preservation
VIII.Critical moments
Figure 17—Events chart showing the relationship of the essential elements and 15. Rift–drift phase
processes, as well as the time of preservation and critical moments of the Permian 16. Late Cretaceous–Cenozoic
petroleum system of Paraná Basin. uplift
3. The intense meteoric gravitational flux, which pro- tial way in response to time and spatial differences in the
vided a renewal of water in the Botucatu and insertion of the igneous intrusions. This sequential mech-
Pirambóia aquifer approximately 500 times over a anism of local pressure relief transformed the generation
period of 100 m.y. (Araújo et al., 1995), would kitchen into a mosaic composed of several different sepa-
biodegrade liquid hydrocarbons moving toward rate and unsynchronized kitchens. It can be inferred that
the gravitational flow, decreasing API gravity and primary migration was lateral, traveling short distances
increasing oil viscosity. Basically, all reservoirs that (1–10 km) to the bordering faults of the compartments,
overlie the Permian source rocks belong to an which were the conduits for upward movement toward
open hydrodynamic environment (Araújo, 1989), the reservoirs.
with an intense flux of meteoric water that has The contemporaneity of the events of generation, pri-
been flushing the porous sandstones since the mary migration, and secondary migration promoted
uplift of the basin margin. multiple generation–migration pulses from different
kitchens. Expulsion efficiency was high due to the cre-
ation of a network of horizontal fractures that intercon-
EVENTS CHART nected the abnormally pressurized kitchens with the
faults that limited each block.
The events chart (Figure 17) summarizes the tempo- Through this vertical fault-focused migration, the
ral relationships of the essential elements and processes petroleum charge reached the carrier beds between 100
of this atypical Permian petroleum system. Due to the and 200 m above the source rock. After passing the Serra
intrusion of diabase sills from a few centimeters up to Alta Formation shaly section, the fragmented petroleum
240 m in thickness during the Cretaceous magmatic charge gradually migrated laterally into thin bodies of
event, high levels of thermal alteration affected the fine sandstone and grainstone of the Teresina Formation
Permian source rocks (average thickness 19 m) that and into thick bodies of fluvial-eolian sandstone of the
reached the liquid and gaseous hydrocarbon generation Rio do Rasto, Pirambóia, and Botucatu Formations (II in
windows (VI in Figure 17). Figure 17).
The magmatic event that occurred between 127 and Temporary and spatial interaction of the migration
138 Ma (Turner et al., 1994; Milani, 1997) produced the and accumulation processes of this atypical Permian
reactivation of megastructures and a complex secondary petroleum system occurred during two phases of trap
network of faults that compartmentalized the petroleum creation. The first was during the Permian–Triassic
system into small blocks (1–10 km2 in area). The activa- Cape–La Ventana orogeny, close to the southern
tion of each compartment probably occurred in a sequen- Gondwana margin (III, 5 and 6, in Figure 17). The second
400 Araújo et al.
phase occurred during Gondwana rifting, concomitant to created by the diachronism of several intrusive pulses
the generation–migration process (magmatic event), with a large range in thickness. The presence of thou-
when transtensional stresses affecting the Jurassic– sands of compartments recorded in the seismic sections
Triassic section (Pirambóia and Botucatu Formations) indicates short distances of primary lateral migration
overlapped a structural framework previously imprinted from the intrusive bodies, estimated to be from 1 to 10 km
during the Permian–Triassic (III, 7, in Figure 17). in distance.
The critical moments happened during the magmatic Secondary migration occurred mostly vertically
event most likely due to diachronism between the cre- through the compartment boundary faults up to the car-
ation of structural inversion (trap formation) and the sec- rier beds located 100–200 m above the Irati source rocks.
ondary migration pulses (VIII, 15, in Figure 17). A second The migration of the charge within the carrier beds is
phase of critical moments occurred during the uplift of interpreted to have occurred along short distances due to
the Paraná Basin margins related to South Atlantic rifting the irregular relief created by the differential intrusion
and the Andean orogeny, which climaxed about 100 and thicknesses, the fragmentation of the charge in thousands
90 Ma (Lelarge, 1993). The uplift caused the erosion of the of primary pulses, and the intense tectonic compartmen-
volcanic Serra Geral sequence, reaching the basal forma- talization caused by the igneous intrusions.
tions of the sedimentary column. Consequently, a gravi- The uplift of the margins of the basin, which occurred
tational hydraulic head was developed concomitant to about 100 Ma, caused the remigration of hydrocarbons
the beginning of tertiary migration (VIII, 16, in Figure 17). from possible traps with consequent loss of a substantial
The resulting intense meteoric gravitational flux, with a portion of the original charge.
renewal of water about 500 times over a period of 100
m.y. (Araújo et al., 1995), constitutes a critical moment for
the integrity of the petroleum system.
Acknowledgments—The authors would like to thank
Petrobrás, Rio de Janeiro, for permission to publish this paper,
Márcio Rocha Mello for the invitation to contribute to this
CONCLUSIONS Memoir, and Luiz Fernando De Ros for an early review of the
manuscript. Additional thanks to Eugênio V. dos Santos Neto
The Irati source rocks, having a large range of thick- for reviews and suggestions on the text and figures and to Paulo
nesses (from a few centimeters up to 240 m), were intrud- R. Veloso for transforming all colored figures into black and
ed by diabase sills related to Serra Geral magmatism white.
(138–127 Ma). The thermal effect extends above and
below the intrusions in a thickness equivalent to that of
the intrusive bodies. In the thermally affected zones,
hydrogen indices decay to 10% and TOC is depleted to REFERENCES CITED
about 30% of its original value.
Al-Shaieb, Z., J. O. Puckette, A. A. Abdalla, and P. B. Ely,
Mass–balance calculations indicate that, as a conse- 1994. Three levels of compartmentation within the over-
quence of this intrusion-driven generation, 500–3500 × pressured interval of the Anadarko Basin, in P. J. Ortoleva,
103 m3 HC/km2 (equivalent oil) of hydrocarbons migrat- ed., Basin compartments and seals: AAPG Memoir 61,
ed out of the source rock in 50% of the Paraná Basin. The p. 69–85.
homogeneization of the temperatures measured from oil Araújo, L. M., 1989. Ambientes hidrodinâmicos da Bacia do
fluid inclusions in calcite and quartz veins in the intru- Paraná—análise preliminar: Internal report,
sion-affected zone range from 60° to 260°C, indicating a Petrobrás/DEPEX/NEXPAR, Curitiba, 23 p.
large range of capture temperature. This probably result- Araújo, L. M., A. B. França, and P. E. Potter, 1995, Aqüífero
ed from the difference between the maximum level of gigante do Mercosul no Brasil, Argentina, Paraguai e
Uruguai: mapas hidrogeológicos das Formações Botucatu,
heat in the source rocks and the cooling that occurred Pirambóia, Rosário do Sul, Buena Vista, Misiones e
during migration, as well as from the reverse migrations Tacuarembó: UFPR/Petrobrás, Curitiba, 16 p.
of less thermally evolved hydrocarbons toward the Bradley, J. S., and D. E. Powley, 1994, Pressure compartments
intrusions. in sedimentary basins: a review, in P. J. Ortoleva, ed., Basin
The high efficiency of primary migration verified in compartments and seals: AAPG Memoir 61, p. 3–27.
the Irati thermally altered zones can be attributed to two Cerqueira, J. R., and E. V. Santos Neto, 1986, Projeto Análise
factors: (1) the presence of a network of predominantly da Bacia do Paraná (Geoquímica Orgânica): Internal
horizontal microfractures and fractures detected in the report, Petrobrás/CENPES/SINTEP, Rio de Janeiro, 3 vol.
examination of thin sections, and (2) the extensive region- Cerqueira, J. R., and E. V. Santos Neto, 1990, Caracterização
al and local tectonic compartmentalization that allowed geoquímica das rochas geradoras de petróleo da
Formação Irati e dos óleos a ela relacionados, Bacia do
the simultaneous occurrence of both primary and sec- Paraná, Brasil: II Congresso Latinoamericano de
ondary migration. Geoquímica Orgânica, Caracas, Venezuela, p. 26.
Large-scale block buoyancy produced by the large vol- Clayton, J. L., and N. H. Bostick, 1986, Temperature effects on
ume of intrusions generated complex differential relief kerogen and on molecular and isotopic composition of the
causing a convergent flow path. The produced charge organic matter in Pierre Shale near an igneous dike:
was fragmented due the presence of several oil kitchens Organic Geochemistry, v. 10, p. 135–143.
Chapter 26—Atypical Permian Petroleum System of Paraná Basin, Brazil 401
Cooles, G. P., A. S. Mackenzie, and T. M. Quigley, 1986, Larter, S., and I. Horstad, 1992, Migration of hydrocarbon
Calculation of petroleum masses generated and expelled into Brent Group reservoirs: some observations from the
from source rocks: Organic Geochemistry, v. 10, Gullfaks Field, Tampen Spur area, North Sea, in A. C.
p. 235–245. Morton, R. S. Haszeldine, M. R. Giles, and S. Brown, eds.,
Demaison, G., and B. J. Huizinga, 1991, Genetic classification Geology of the Brent Group: Geological Society of London
of petroleum systems: AAPG Bulletin, v. 75, p. 1626–1643. Special Publication, v. 61, p. 441–452.
Dennis, L. W., G. E. Maciel, P. G. Hatcher, and B. R. T. Lelarge, N. L. M. V., 1993, Thermochronologie par la method
Simoneit, 1982, 13C nuclear magnetic resonance studies of des traces de fission d’une marg passive (Dome de Ponta
kerogen from Cretaceous black shales thermally altered by Grossa, SE, Brasil): Thesis de docteur de l’Universit Joseph
basaltic intrusion and laboratory simulations: Geochimica Fourier, Grenoble, France, 244 p.
et Cosmochimica Acta, v. 43, p. 901–907. Macgregor, D. S., and A. S. Mackenzie, 1987, Quantification
Deroo, G., T. G, Powell, B. Tissot, and R. G. McCrossan, 1977, of oil generation and migration in the Malaca Strait
The origin and migration of petroleum in the Western region: Proceedings of 15th Annual Convention of
Canadian sedimentary basin, Alberta; a geochemical and Indonesian Petroleum Association, Jakarta, p. 305–309.
thermal maturation study: Geological Survey of Canada Mackenzie, A. S., and T. M. Quigley, 1988, Principles of geo-
Bulletin, v. 262, 136 p. chemical prospect appraisal: AAPG Bulletin, v. 72, p.
De Wit, M. J., and I. D. Ransome, 1992, Regional inversion 399–415.
tectonics along the Southern margin of Gondwana, in M. Mackenzie, A. S., I. Price, D. Leythaeuser, P. Muller, M.
J. De Wit and I. D. Ransome, eds., Inversion tectonics of Radke, and R. G. Schaefer, 1987, The expulsion of petrole-
the cape fold belt, Karoo and Cretaceous basins of um from Kimmeridge clay source-rocks in the area of Brae
Southern Africa. Rotterdan, Balkema, p. 15–21. field, UK continental shelf, in J. Brooks and K. W. Glennie,
England, W. A., 1994, Secondary migration and accumulation eds., Petroleum geology of north western Europe:
of hydrocarbons, in L. B. Magoon and W. G. Dow, eds., London, Graham and Trotman, p. 865–877.
The Petroleum system—from source to trap: AAPG Magoon, L. B., and W. G. Dow, 1994, The petroleum system,
Memoir 60, p. 211–217. in L. B. Magoon and W. G. Dow, eds., The petroleum sys-
England, W. A., A. S. Mackenzie, D. M. Mann, and T. M. tem—from source to trap: AAPG Memoir 60, p. 3–24.
Quigley, 1987, The movement and entrapment of petrole- Marques, A., O. A. Zanotto, O. B. Paula, M. A. M. Astolfi, A.
um fluids in the subsurface: Journal of the Geological B. França, and E. A. Barbosa, 1993, Compartimentação tec-
Society, v. 144, p. 327–347. tônica da Bacia do Paraná: Internal Report,
England, W. A., A. L. Mann, and D. M. Mann, 1991, Petrobrás/DEPEX/NEXPAR, Curitiba, 26 p.
Migration from source to trap, in R. K. Merrill, ed., Source Mendonça Filho, J. G.,1994, Estudo petrográfico e organo-
and migration processes and evaluation techniques: geoquímico de amostras de folhelhos da Formação Irati,
AAPG Treatise of Petroleum Geology, p. 23–46. Permiano Superior da Bacia do Paraná, no Estado do Rio
Espitalié, J., F. Marquis, and L. Sage, 1987, Organic geochem- Grande do Sul: Tese de Mestrado, IG-UFRGS, Porto
istry of the Paris Basin, in J. Brooks and K. W. Glennie, Alegre, 212 p.
eds., Petroleum geology of north west Europe: London, Milani, E. J., 1992, Intraplate tectonics and the evolution of
Graham and Trotman, v. 1, p. 71–86. the Paraná Basin, SE Brazil, in M. J. De Wit and I. D.
Ferreira J. C., J. R. Cerqueira, and E. S. T. Frota, 1994, Processo Ransome, eds., Inversion tectonics of the cape fold belt,
não convencional de geração e migração de petróleo e Karoo and Cretaceous basins of southern Africa:
suas implicações exploratórias: Internal Report, Rotterdam, Balkema, p. 101–108.
Petrobrás/CENPES/DIVEX/SEGEQ, Rio de Janeiro, 66 p. Milani, E. J., 1997, Evolução Tecto-Estratigráfica da Bacia do
Fuzikawa, K., 1994, Microscopia e medidas das temperaturas Paraná e seu Relacionamento com a Geodinâmica
de homogeneização das inclusões fluidas em vênulas de Fanerozóica do Gondwana Sul-Ocidental: Tese de
carbonatos das amostras de sedimentos da Bacia do doutorado em Geociências, Instituto de Geociências,
Paraná, CNEM, Belo Horizonte, 12 p. Universidade Federal do Rio Grande do Sul, RS, Porto
Gaarenstroom, L., R. A. J. Tromp, M. C. de Jong, and A. M. Alegre, 255 p.
Brandenburg, 1993, Overpressure in the Central North Ortoleva, P. J, 1994, Basin compartments and seals: AAPG
Sea: implications for trap integrity and drilling, in J. R. Memoir 61, 477 p.
Parker, ed., Petroleum geology of northwest Europe: Pepper, A. S., 1991, Estimating the petroleum expulsion
Proceedings of the 4th Conference, Petroleum Geology 86, behaviour of source rocks: a novel quantitative approach,
The Geological Society of London, p. 1305–1313. in W. A. England and A. J. Fleet, eds., Petroleum migra-
Hettich, M., 1981, Arenitos oleígenos da Formação Pirambóia, tion: Geological Society Special Publication 59, p. 9–31.
São Paulo—Segunda Parte: cadastramento regional de Peters, K .E., J. K. Whelan, J. M. Hunt, and M. E. Tarafa, 1983,
ocorrências: Internal Report, GEOSOL/SIX, São Mateus Programmed pyrolysis of organic matter from thermally
do Sul, 133 p. altered Cretaceous black shales: AAPG Bulletin, v. 67, p.
Hunt, J. M., 1990, Generation and migration of petroleum 2137–2146.
from abnormally pressured fluid compartments: AAPG Raymond, A. C., and D. G. Murchison, 1992, Effect of igneous
Bulletin, v. 74, p. 1–-12. activity on molecular-maturation indices in different types
Lafargue, E., J. Espitalié, T. M. Broks, and B. Nyland, 1994, of organic matter: Organic Geochemistry, v. 18, p. 725–735.
Experimental simulation of primary migration: Organic Sad, J. H. G., and N. T. A. Saraiva, 1982, Rochas oleígenas da
Geochemistry, v. 22, p. 575–686. Formação Irati: Internal Report, Geosol-Geologia e
Larter, S., 1988, Some pragmatic perspectives in source rock Sondagens Ltda e Superintendência da Industrialização
geochemistry: Marine and Petroleum Geology, v. 5, do Xisto.
p. 194–204.
402 Araújo et al.
Saxby, J. D., and L. C. Stephenson, 1987, Effect of an igneous Sokolov, V. A., A. A. Geodekyan, C. G. Grigoryev, A. Ya.
intrusion on oil shale at Rundle (Australia): Chemical Krems, V. A. Stroganov, L. M. Zorkin, M. I. Zeidelson, and
Geology, v. 63, p. 1–16. S. J. Vainbaum, 1971, The new methods of gas surveys,
Shepherd, L. D., P. A. Drzewieckie, J. M. Bahr, and J. A.Simo, gas investigations of wells and some practical results, in
1994, Silica budget for a diagenetic seal, in P. J. Ortoleva, R. W. Boyle, ed., Geochemical exploration: Canadian
ed., Basin compartments and seals: AAPG Memoir 61, Institute of Mining and Metallurgy Special Volume 11,
p. 369–385. p. 538–544.
Silva, R. B. G., 1983, Estudo hidroquímico e isotópico das Surdan, R. C., Z. S. Jiao, and R. S. Martinsen, 1994, The
águas subterrâneas do aqüífero Botucatu no Estado de São regional pressure regime in Cretaceous sandstones and
Paulo: Tese de Doutorado, USP, São Paulo, 133 p. shales in the Powder River Basin, in P. J. Ortoleva, ed.,
Simoneit, B. R. T., S. Brenner, K. E. Peters, and I. R. Kaplan, Basin compartments and seals: AAPG Memoir 61,
1978, Thermal alteration of Cretaceous black shale by p. 213–235.
basaltic intrusions in the eastern Atlantic: Nature, v. 273, Trigüis, J. A., 1986, An organic geochemistry investigation of
p. 501–504. heat-affected sediments in the Paraná Basin (Brasil): Ph.D.
Simoneit, B. R. T., S. Brenner, K. E. Peters, and I. R. Kaplan, Thesis, University of Newcastle, Newcastle, U.K., 203 p.
1981, Thermal alteration of Cretaceous black shale by dia- Turner, S., M. Regelous, S. Kelley, C. Hawkesworth, and M.
base intrusions in the Eastern Atlantic—II. Effects on bitu- Mantovani, 1994, Magmatism and continental break-up in
men and kerogen: Geochimical et Cosmochimica Acta, the South Atlantic: high precision Ar/Ar geochronology:
v. 45, p. 1581–1602. Earth and Planetary Science Letters, Amsterdam, v. 121,
Sluijk, D., and M. H. Nederlof, 1984. Worldwide geological p. 333–348.
experience as a systematic basis for prospect appraisal, in Zanotto, O. A., 1993, Erosão pós-Cretáceo na Bacia do Paraná,
G. Demaison and R. J. Murris, eds., Petroleum geochem- com base em dados de reflectância da vitrinita: Simpósio
istry and basin evaluation: AAPG Memoir 35, p. 15–26. Sul-brasileiro de Geologia, 5, Sociedade Brasileira de
Geologia, Resumos, Curitiba, p. 58.