chap10

94
Chapter 10 Evaluating Top and Fault Seal by Grant M. Skerlec

Upload: sagar-kumar

Post on 14-May-2015

734 views

Category:

Education


6 download

TRANSCRIPT

Page 1: Chap10

Chapter 10

Evaluating Top and Fault Seal

by

Grant M. Skerlec

Page 2: Chap10

Grant M. SkerlecGrant M. Skerlec is head of SEALS International, specializing in the application of faultseal analysis to prospect assessment and field development. He received a BA degreefrom Franklin & Marshall College in 1968 and a Ph.D. from Princeton University in1979. He joined Exxon Production Research Company where he began working on faultseal analysis in 1978. In 1982 he joined Esso Norway, and later Esso Exploration &Production UK, as an exploration geologist applying quantitative seal analysis toprospect and play assessments in the North Sea, Norwegian Sea, and Barents Sea. In1987 Skerlec formed SEALS International, a company that provides services, software,and training for routine fault seal analysis as well as global databases of fault seal behav-ior in hydrocarbon basins. Current work involves the compilation of the joint-industryFault Seal Behavior in the Gulf Coast Atlas and the recent completion of a similar atlasfor North Sea–Norwegian Sea oil and gas fields.

Page 3: Chap10

Overview • 10-3

This chapter discusses seal analysis techniques. Both top seal and fault seal are funda-mental to prospect and play assessment as well as to production and field development.Despite our understanding of the variables that control seals (Downey, 1984), practicaltechniques are few and seal is commonly risked in an intuitive, qualitative manner.However, quantitative seal analysis, using those few techniques available, improves suc-cess ratios and reduces costly errors in field development.

Introduction

Overview

Top seal and fault seal are important because they control the following:• Presence or absence of hydrocarbons• Percent fill• Vertical and lateral distribution of hydrocarbons• Migration pathways and charge volumes• Distribution and movement of hydrocarbons during field development

Seals are fundamental; no seal, no trap. Seals, or their absence, also define leak pointsthat control the percent fill for hydrocarbon accumulations. Assessment of percent fillwithout the ability to risk seal (top or fault) is reduced to a statistical guessing game.They control the vertical and lateral distribution of hydrocarbons, both within individualfields and within basins.

Seals control migration pathways into traps. A trap may be empty not because a faultleaked once-trapped hydrocarbons but because a fault sealed and prevented hydrocarbonsfrom migrating into a trap and filling it in the first place. Similarly, top seals can restrictvertical migration into shallow traps and control the vertical and lateral distribution ofhydrocarbons within a basin.

Hydrocarbons migrate until they encounter the first intact seal. Because of variations inseal integrity and capacity, drainage areas are four-dimensional. Prospect analysis usingdrainage areas defined by simple structure-depth maps on the top reservoir can be verymisleading. Plays appear and disappear in response to seal behavior.

Seals also control the movement of hydrocarbons during production. Efficient field devel-opment, well placement, ultimate recovery, and economic success or failure depend onrisking seal.

Importance ofseal

This chapter contains the following sections.

Section Topic Page

A Evaluating Fault Seal 10–4

B Evaluating Top Seal Integrity 10–45

C Evaluating Intact Top Seal 10–64

D References 10–88

In this chapter

Page 4: Chap10

10-4 • Evaluating Top and Fault Seal

Routine fault seal analysis of prospects and fields is a necessary part of all explorationand production. The sealing behavior of faults is important because it controls the follow-ing:• Trapped hydrocarbon volumes—faults control the presence or absence of hydrocar-

bons, the percent fill of individual fault compartments, and the vertical and lateral dis-tribution of hydrocarbons within a field.

• Migration pathways into traps—faults control both the volume and migrationdirection of hydrocarbons available to charge a trap.

• Hydrocarbon movement during field development—faults can retard hydrocar-bon flow, limit sweep efficiency, and create isolated compartments bounded by sealingfaults. If the faults leak, they provide field-wide communication among numerous faultcompartments.

Introduction

Section A

Evaluating Fault Seal

Evaluating, or risking, fault seal is a four-step process, as outlined in the table below anddiscussed in detail within this section.

Step Task

1 Establish empirical threshold for seal/leak in existing fields.

2 Construct a fault plane profile to show juxtaposition relationships.

3 Analyze fault seal quantitatively to predict seal behavior and identifyfault-dependent leak points.

4 Construct migration pathway maps.

The riskingprocess

This section contains the following subsections.

Subsection Topic Page

A1 Fault Seal Behavior 10–5

A2 Fault Plane Profile Analysis 10–17

A3 Quantitative Fault Seal Analysis 10–22

A4 Fault Seal and Migration Pathways 10–35

A5 How Faults Affect Field Development 10–40

In this section

Page 5: Chap10

Evaluating Fault Seal • 10-5

Understanding fault seal behavior in existing fields is a prerequisite for predicting sealbehavior in untested prospects. Empirical studies have established patterns of sealbehavior, identified seal/leak thresholds for quantitative analysis of a fault seal, providedreal analogs for prospect assessment, and demonstrated how important fault seal behav-ior is in controlling hydrocarbon accumulations (Smith, 1980; Skerlec, 1990, 1997a,b;Yielding et al., 1997).

Introduction

Subsection A1

Fault Seal Behavior

This subsection contains the following topics.

Topic Page

Fault Seal Behavior Basics 10–6

Cross-leaking Faults 10–7

Cross-sealing Faults 10–10

Dip-sealing Faults 10–12

Dip-leaking Faults 10–13

Controls on Percent Fill 10–16

In thissubsection

Page 6: Chap10

10-6 • Evaluating Top and Fault Seal

Fault seal behavior is analyzed on the basis of the following:• Hydrocarbon contacts• Fault-dependent leak points• Pressure data

Introduction

Fault Seal Behavior Basics

The two basic types of fault seal behavior are (1) cross sealing or cross leaking and (2) dipsealing or dip leaking. Cross seal and leak refers to the lateral communication acrossthe fault between juxtaposed sands. Dip seal and leak refers to the vertical communica-tion along the fault between stacked sands.

The type of seal behavior is important in controlling the type of fault-dependent leakpoints. Fault-dependent leak points limit the volume of trapped hydrocarbons. The abilityto identify these fault-dependent leak points is a fundamental tool for prospect assess-ment.

Types ofbehavior

Faults may have some finite seal capacity. A fault might support the pressure exerted bya 50-m hydrocarbon column yet leak if the column increases to 51 m (Smith, 1980).

Finite sealcapacity

A fault does not simply seal or leak; many variations exist:• A fault can seal at one point and leak at another. • A fault can juxtapose many individual reservoirs, and each fault/reservoir intersection

can seal or leak independently. • A fault can seal on one side and leak on the other. • A fault can seal oil but leak gas. • A fault can be sealing to some finite column of hydrocarbon but leaking to a larger

column. • A fault can change seal behavior during migration and fill as well as during produc-

tion. Seal behavior is time dependent.

Caveat

Page 7: Chap10

Evaluating Fault Seal • 10-7

A cross-leaking fault allows lateral communication of hydrocarbons between juxtaposedreservoirs. Cross-leaking faults can be identified using any of the following criteria:• Common hydrocarbon contacts• Common free water levels (FWL)• Juxtaposed lithology leak points (JLLP)• Common pressures

Introduction

Cross-leaking Faults

Common hydrocarbon contacts imply communication across the fault and cross leakage.The cross-leaking fault in the following figure shows two sands, Ru and Rd, juxtaposed bya fault. The two sands have common oil–water (OWC) and gas–water (GWC) contacts.The fault is cross leaking to both oil and gas.

Commonhydrocarboncontacts

Figure 10–1.

A cross-leaking fault can have different hydrocarbon contacts across the fault. The differ-ence in hydrocarbon contacts can be caused not by the fault zone material but by differ-ences in the displacement pressure (Pd) of the juxtaposed reservoirs. There is, however, acommon free-water level (FWL).

Common free-water levels

An example of a cross-leaking fault with different OWCs and a common FWL is shown inthe following figure. The Pd of the Rd sand is greater than that of the Ru sand. The fault iscross leaking despite different OWCs.

Figure 10–2.

Page 8: Chap10

Cross leakage commonly creates fault-dependent leak points limiting the percent (Smith,1966, 1980; Allan, 1989; Harding and Tuminas, 1989). One type of fault-dependent leakpoint is illustrated in the following figure. The coincidence of the hydrocarbon contactwith the top of the sand juxtaposed across the fault is a juxtaposed lithology leak point(JLLP). Hydrocarbons are trapped only where there is sand/sand juxtaposition along thefault. Hydrocarbons leak across the sand/sand juxtapositions.

Juxtaposedlithology leakpoints

10-8 • Evaluating Top and Fault Seal

Cross-leaking Faults, continued

The following figure illustrates the effect of capillary properties on oil–water contacts.Decreasing pore throat radius, represented by three capillary tubes of decreasing diame-ter (left), creates a higher OWC within the reservoir. If the pore throat is large (low Pd),the OWC coincides with the free water level. If the pore throat is small (high Pd), theOWC is higher than the free water level. In a reservoir with a lateral facies change, afault can be cross leaking but still separate sands with different hydrocarbon contacts(right).

Capillarity andOWCs

Figure 10–3.

Figure 10–4.

Identifying JLLPs is an important method of assessing percent fill in prospects and deter-mining seal behavior in existing fields. JLLPs exist only if the fault is cross leaking.

Page 9: Chap10

Evaluating Fault Seal • 10-9

Cross-leaking Faults, continued

Common pressures across a fault imply communication and cross leakage. If a new wellin a separate fault compartment encounters pressures equal to the current field depletedpressures, the fault is cross leaking.

In the figure below, wells 1 and 2 are separated by a cross-leaking fault. The initial pres-sures of both wells lie on a common, field-wide pressure depletion curve.

Commonpressures

Figure 10–5.

In fields with a long, complex production history, pressures and hydrocarbon contacts canbe misleading as indicators of long-term fault seal behavior. Different pressures can existacross a fault despite cross leakage. Pressures in these fields may reflect short-term, pro-duction-induced disequilibrium. Pressure differences may exist across a fault that wascross leaking during migration and fill; the implied cross seal may not have existed dur-ing migration and fill when the fault was at equilibrium. The longer time spans of migra-tion and fill allow equilibrium; the short spans of production favor disequilibrium.

In addition, the pressure distribution within a closure is affected by all of the adjacentfaults, the rates of production and depletion of individual fault compartments, reservoirpermeability and continuity, and the relative permeability of both reservoirs and faultzones. Apparent "permeability barriers" within a fault compartment may also be artificialcreations of more distant bounding faults (van Poollen, 1965; Prasad, 1975; Earlougherand Kazemi, 1980; Stewart et al., 1984; Yaxley, 1987).

Caveat

Page 10: Chap10

10-10 • Evaluating Top and Fault Seal

A cross-sealing fault prevents communication of hydrocarbons between juxtaposed sands(reservoirs). Cross-sealing faults can be identified using the following criteria:• Hydrocarbon-bearing sands against water-wet sands• Different hydrocarbon contacts• Different pressures

Introduction

Cross-sealing Faults

A fault is cross sealing if it juxtaposes hydrocarbon-bearing sands with water-wet sands,as illustrated in the following figure. Both oil and gas are prevented from flowing into thesand in the hanging wall (Rd) by the cross-sealing fault.

Hydrocarbonagainst water

Figure 10–6.

A fault is also cross sealing if it juxtaposes sands with different hydrocarbon contactsand/or different free water levels, as illustrated in the following figure. Small differencesin hydrocarbon contacts do not necessarily imply a cross-sealing fault because the capil-lary properties of the juxtaposed sands can create different hydrocarbon contacts evenacross a cross-leaking fault. Different free water levels do imply a cross-sealing fault.

Differenthydrocarboncontacts

Figure 10–7.

Page 11: Chap10

Evaluating Fault Seal • 10-11

Different pressures across a fault imply cross seal. In the figure below, wells 1 and 2 areseparated by a cross-sealing fault. Well 2 encountered virgin pressures in contrast to thelower pressures in the main field.

Differentpressures

Cross-sealing Faults, continued

Figure 10–8.

Page 12: Chap10

10-12 • Evaluating Top and Fault Seal

Introduction

Dip-sealing Faults

Figure 10–9.

Dip-sealing faults are important because they can create purely fault-dependent traps.No independent structural closure is required for entrapment. Where independent struc-tural closure does exist, as in the figure below, a dip-sealing fault can trap additional vol-umes of oil against the fault. Dip-sealing faults can trap hundreds of meters of oil withoutindependent closure. In the following figure, both oil and gas are trapped against the faultand have not leaked up the fault zone.

A dip-sealing fault traps hydrocarbons against the fault plane.

Importance ofdip-sealing faults

Page 13: Chap10

Evaluating Fault Seal • 10-13

A fault is dip leaking if hydrocarbons have migrated along the fault plane. Dip-leakingfaults can be identified by the presence of a fault plane leak point (FPLP).

Introduction

Dip-leaking Faults

An FPLP is a type of fault-dependent leak point in which the hydrocarbon contact coin-cides with the intersection of the fault plane and the top of the reservoir. As shown in thefollowing figure, an FPLP limits the hydrocarbon to the structurally independent closure.The lack of hydrocarbons in contact with the fault plane implies leakage has occurred ver-tically along the fault (Smith, 1966; Allard, 1993; Harding and Tuminas, 1988).

Fault planeleak point(FPLP)

Figure 10–10.

The ability to predict leakage is important in prospect assessment. An FPLP limits thehydrocarbon volume. Where no independent closure exists, prospects may be completelyemptied by dip leakage.

A fault may dip seal on one side of a fault and dip leak on the other. This asymmetric dipleakage is caused primarily by variations in the sand–shale ratio of the fault gouge. Otherpossible controls in some basins include asymmetric fracture density and/or preferentialhydraulic fracturing in the hanging wall (Weber et al., 1978; Skerlec, 1990).

Asymmetric dipleakage

Page 14: Chap10

10-14 • Evaluating Top and Fault Seal

An example of asymmetric dip leakage is the Chocolate Bayou field, U.S. Gulf Coast,shown in the following figure. All of the hanging wall gas accumulations are limited byFPLPs and dip leak. The three gas accumulations in the footwall, however, are all dipsealing; all have gas columns in contact with the fault plane. This pattern of behavior iscommon in both the U.S. Gulf Coast and the Niger Delta (Weber et al., 1978).

Example ofasymmetric dipleakage

Dip-leaking Faults, continued

Figure 10–11. After Myers, 1968; courtesy Gulf Coast Assoc. of Geologists.

Low-side traps may have very different fault seal risk compared with high-side traps. Inthe following figure, a map shows that dip leakage in the hanging wall limits fill (shaded)to the structurally independent closure. Dip seal in the footwall allows fill in excess of theindependent closure. The footwall contains purely fault-dependent traps. The hangingwall relies upon independent closure for entrapment.

Low-side traps

Figure 10–12.

Page 15: Chap10

Evaluating Fault Seal • 10-15

A fault may dip seal oil and dip leak gas or vice versa. The following illustration shows afault that is dip leaking to gas in both the hanging wall and footwall. Both gas accumula-tions are limited by an FPLP. The oil accumulation in the footwall, however, is filledbeyond the structurally independent closure. The fault is dip sealing to oil but dip leakingto gas.

Dip leakage ofgas vs. oil

Dip-leaking Faults, continued

Figure 10–13.

Although dip leakage is generally understood as the migration up the fault zone due tobuoyant forces, leakage down the fault zone can also occur. Refraction of flow lines canoccur when fluids cross a fault zone of differing permeability from the wall rock. Fluidflow across a fault of lower permeability than the adjacent wall rock could dip leak downthe fault zone (Hubbert, 1953). Similarly, a fault zone with higher permeability than thewall rock will refract flow lines up the fault zone. Pressure differences along and acrossfaults can also cause down-dip flow in the area of lowest pressure (Knutson and Erga,1991; Niemann and Krowlow, 1992).

Direction of dipleakage

Faults in most basins are at equilibrium. Those faults that were going to leak haveleaked. The present structure and juxtapositions control fault seal behavior and spillpoints. However, disequilibrium may exist in basins now undergoing rapid migration andfill. The Los Angeles basin (California) is one example. Faults may be dip leaking, but therate of charge is greater than the rate of leakage. These faults appear to dip seal—despiteseeps along the surface trace of the faults, which indicate dip leakage.

Caveat

Page 16: Chap10

10-16 • Evaluating Top and Fault Seal

The percent fill of a trap is the percentage of the trap volume filled with hydrocarbonscompared with its total volume. Percent fill can be controlled by a number of factors,including the following:• Top seal capacity/integrity• Synclinal spill points• Charge• Fault seal capacity• Fault-dependent leak points

Introduction

Controls on Percent Fill

The following figure summarizes the various controls on percent fill and hydrocarbon col-umn height. The maximum fill case is the synclinal spill point (SSP). A cross-leaking faultlimits the percent fill to the juxtaposed lithology leak point (JLLP). A dip-leaking faultlimits the percent fill to the fault plane leak point (FPLP).

A cross-sealing fault allows fill below the JLLP and possibly as deep as the SSP. A trapalso may be partially filled (PF) due to either charge or top seal capacity. A partially filledtrap can have a hydrocarbon contact at any depth.

Controlsummary

Figure 10–14.

Sealing faults may have some finite seal capacity. A fault may be able to support the pres-sure exerted by a 50-m hydrocarbon column but leak if the column increases to 51 m(Smith, 1980). Recent work suggests there may be a maximum seal capacity related tothe percentage of shale in the fault gouge (Yielding et al., 1997; Skerlec, 1997b). In mostcases, however, the percent fill is significantly less and is limited by fault-dependent leakpoints rather than the seal capacity of the gouge.

Finite sealcapacity offaults

Page 17: Chap10

Evaluating Fault Seal • 10-17

A fault plane profile is a cross section in the plane of the fault that shows both the hang-ing wall and footwall cutoffs (Van Wijhe et al., 1980; Allan, 1989; Harding and Tuminas,1989). Fault plane profiles are a fundamental tool for prospect assessment as well as afirst step in understanding seal behavior in existing fields. Fault plane profiles are impor-tant because they show what is being juxtaposed across the fault. By doing so, they showareas of sand/sand and sand/shale juxtaposition, establish seal relationships, definepotential fault-dependent leak points, and help assess hydrocarbon volumes.

Introduction

Subsection A2

Fault Plane Profile Analysis

This subsection contains the following topics.

Topic Page

Constructing a Fault Plane Profile 10–18

Avoiding Pitfalls of Fault Plane Profile Analysis 10–20

In thissubsection

Page 18: Chap10

10-18 • Evaluating Top and Fault Seal

The fault cross-leaks along its entire surface at sand/sand juxtapositions. At eachsand/sand juxtaposition, a JLLP spills the hydrocarbons across the fault. There is suffi-cient charge to fill all closures to a fault-dependent leak point. Potential reservoirs withJLLPs at the crest are dry. The percent fill for other sands is constrained by the JLLPs(Allan, 1989). If bed dips in the hanging wall allow entrapment against the fault, hydro-carbons could be trapped in both the hanging wall and footwall reservoirs. With sufficientcharge, common hydrocarbon contacts could exist.

A fault plane profile (FPP) is a cross section in the plane of the fault that shows both thehanging wall and footwall cutoffs.

An FPP is distinctive in two ways:1. It is a cross section in the plane of the fault rather than normal to the fault.2. Both the hanging wall and footwall stratigraphic cutoffs are shown on the same cross

section.

Introduction

Constructing a Fault Plane Profile

We construct an FPP as we do any geologic cross section. Using structure–depth maps, weproject the top of each mapped stratigraphic unit at its intersections with the fault planeto its correct depth on the fault plane profile (Allan, 1989).

Because only a fraction of the stratigraphy is mapped routinely, well logs are used to pro-ject the remainder of the detailed stratigraphy onto the FPP.

Procedure

The following figure is a simple example of a fault plane profile. The sand reservoirs inthe hanging wall (dark gray) and footwall (light gray) are shown on the same cross sectiondrawn in the plane of the fault. In this example, all hydrocarbon accumulations (black)are limited by JLLPs (juxtaposed lithology leak points). Hydrocarbons cross-leak from thefootwall sands into the hanging wall sands.

Example: FPPof cross-leakingfault

Figure 10–15.

Page 19: Chap10

Evaluating Fault Seal • 10-19

Constructing a Fault Plane Profile, continued

The fault plane profile in the following figure is similar to the profile in Figure 10–15except that the fault is cross sealing at the fault/reservoir intersection in the lowest foot-wall sand. Despite a sand/sand juxtaposition at the crest of this reservoir, the fault trapshydrocarbons at a cross sealing segment. Higher sands continue to cross-leak. The per-cent fill of the lowest sand is limited by either the Pd of the fault zone, the charge volume,or the top seal—not a JLLP.

Example: FFP that crossseals, crossleaks

Figure 10–16.

For an example of a fault plane profile of a fault in a Gulf Coast field, refer to Figure10–26.

Page 20: Chap10

10-20 • Evaluating Top and Fault Seal

There are three pitfalls to avoid in using fault plane profiles:• Assuming cross leakage• Incorrectly projecting bed thickness and bed tops• Erroneous structural interpretations

Pitfalls

Avoiding Pitfalls of Fault Plane Profile Analysis

We must not assume that a fault always cross-leaks at sand/sand juxtapositions. Crossseal is common. If the fault is cross sealing, it is impossible to identify JLLPs from faultplane profiles. Seal behavior can be predicted using quantitative fault seal analysis (referto Subsection A3).

Assuming crossleakage

If all bed boundaries are taken from mapped surfaces, there are no problems. Becausethis is almost never the case, problems arise when unmapped bed boundaries, includingbed bases, are plotted on fault plane profiles by "isopaching down" from mapped surfaces.Errors arise due to confusion regarding apparent thickness. Errors in bed thickness leadto mistaken identity of what is juxtaposed along the fault as well as mistaken leak pointsand percent fill.

Plotting bedthicknessincorrectly

Fault plane profiles can beconstructed in two differentways:• Constructed in the plane

of the fault• Projected onto a vertical

plane parallel to the faulttrace

The illustration to the rightshows bed cutoffs on thefault plane as well as thoseprojected onto a verticalfault plane profile. The topof the bed is correctly plot-ted at a depth of 1000 ftalong both the fault planeand the vertical profile.

Constructionmethods

Figure 10–17.

Page 21: Chap10

Evaluating Fault Seal • 10-21

Projecting the base of the bed requires determining the correct apparent thickness. Thetrue stratigraphic thickness is H. Several apparent thicknesses are involved:• Hw—apparent thickness measured in a well log• Hf—apparent thickness along the plane of the fault• Hp—apparent thickness along the fault plane profile

Apparentthickness

Avoiding Pitfalls of Fault Plane Profile Analysis, continued

Use of the well-log thickness (Hw) to project the base on either fault plane profile clearlygives an incorrect bed thickness at Hp2 in Figure 10–17. Note that H = Hf only because thebed is normal to the fault plane in this specific case. The correct apparent thickness to beplotted on the fault plane profile is Hp1 at 1100 ft.

Well-logthickness

The correct thickness can be determined from trigonometry (Skerlec, 1990) or from a sim-ple proportion, as illustrated in the following figure. The base of the sand is plotted byusing a simple proportion between the thickness in the well (Hw) and the distance to thenext mapped horizon in the well (Dw), and between the thickness (Hp) on the FPP and thedistance to the next mapped horizon on the FPP (Dp):

Hw/Dw = Hp/Dp

Determiningcorrectthickness

Figure 10–18.

The interpretation of the juxtaposition relationships along the fault plane profile is onlyas good as the original maps. Older 2-D seismic surveys are inadequate unless there isabundant well control. One of the indirect benefits of routine fault seal analysis is that itfocuses the explorationist on the details of fault geometry and fault-bed cutoffs that arecritical to trap definition as well as to seal analysis. Modern 3-D seismic surveys areessential for reliable risk of fault seal. It is also true that using expensive 3-D seismic sur-veys only to make top of reservoir structure maps without fault seal analysis wastes a sig-nificant part of the information available for prospect assessment.

Erroneousstructuralinterpretation

Page 22: Chap10

10-22 • Evaluating Top and Fault Seal

Modern quantitative methods of risking fault seal properties have established an empiri-cal relationship between the observed sealing behavior of faults and the inferredsand–shale content of the fault zone (Bouvier et al., 1989; Allard, 1993; Jev et al., 1993;Skerlec, 1997a,b; Yielding et al., 1997).

Introduction

Subsection A3

Quantitative Fault Seal Analysis

Quantitative fault seal analysis lets us predict fault seal behavior. Fault seal behavior israrely random. Once we understand the pattern of seal behavior in existing fields, we canrisk seal behavior in untested prospects.

Importance ofquantitativeanalysis

This subsection contains the following topics.

Topic Page

How Fault Zones Affect Seal 10–23

Quantitative Fault Seal Analysis 10–26

Limitations of Quantitative Fault Seal Analysis 10–30

Example of Routine Fault Seal Analysis: Gulf Coast 10–31

In thissubsection

Page 23: Chap10

Evaluating Fault Seal • 10-23

Fault zones in siliciclastic sequences consist of various proportions of sand and shale.There is a relationship between seal behavior and the inferred sand–shale ratio of thefault zone. Sand-rich portions of the fault zone leak; shale-rich portions of the fault zoneseal. Given the complex structure and composition of fault zones and the many factorsthat could affect a seal, it has been surprising to find such a simple relationship, basinafter basin.

Introduction

How Fault Zones Affect Seal

The following figure is an example of smear gouge along a normal fault found in a se-quence of Permian–Carboniferous sand and shales near Lynemouth, U.K. Ductile, lightgray shales have been incorporated in the fault zone as a continuous layer sourced from abed in the footwall (upper left). Similarly, a bed of coal (black) in the footwall and sand-stone in the hanging wall have been incorporated as discontinuous fragments in the faultzone. The coal, however, is intensely fractured, reflecting brittle rather than ductile defor-mation.

Both brittle sands and coals as well as ductile shales are incorporated in the fault zone.The mechanism of gouge formation is much more complex than a simple "smearing" ofductile shales into the fault zone. Although faults exhibit a complete spectrum of behaviorfrom brittle to ductile, interbedded sand–shale sequences with strong ductility contrastcommonly exhibit this transitional behavior.

Example:Smear gouge

Figure 10–19.

Page 24: Chap10

10-24 • Evaluating Top and Fault Seal

Shale-rich gouge zones can create impermeable barriers. The following figure, a normalfault in Jurassic and Triassic sediments near Lilstock, U.K., shows a gouge zone approxi-mately 5 m thick. The gouge consists predominantly of ductile shales and marls as well assmall amounts of brittle limestones. Faults with similar shale-rich gouge in producingfields create seals.

Example: Smear gouge(continued)

How Fault Zones Affect Seal, continued

Figure 10–20.

Page 25: Chap10

Evaluating Fault Seal • 10-25

Fault gouge can range from sand rich to shale rich. The following figure shows three faultzones exhibiting a spectrum of sand–shale ratios. The table below describes the positionof the three faults, their sequences and characteristics, and their location.

Position Sequence Characteristics Location

Left Sand–siltstone Gouge with high sand–shale ratio Tertiary, Isle of Wight, U.K.

Center Limestone–shale Gouge with intermediate "sand"–shale ratio Jurassic, Kilve, U.K.

Right Limestone–shale Gouge with low "sand"–shale ratio Jurassic, Kilve, U.K.

Example:Spectrum ofgougecompositions

How Fault Zones Affect Seal, continued

Figure 10–21.

Page 26: Chap10

10-26 • Evaluating Top and Fault Seal

Procedure

The capacity of a fault to leak or seal hydrocarbons is largely controlled by the smear–gouge ratio (SGR). The SGR is an estimate of the composition of that portion of the faultzone through which leakage or seal must occur. We calculate the SGR by measuring thecumulative shale and sand that has moved past that zone. A fault cross-leaks or dip-leaksif the sand–shale ratio is high in the zone (high SGR). A fault cross-seals or dip-seals if thesand–shale ratio is low in the zone (low SGR).

Several other algorithms for estimating gouge composition exist (Bouvier et al., 1989;Allard, 1993; Jev et al., 1993; Gibson, 1994; Yielding et al., 1997). All have established arelationship between the actual seal behavior and inferred gouge composition.

Introduction

Quantitative Fault Seal Analysis

The following table outlines the procedure for quantitative fault seal analysis.

Step Task

1 Analyze logs to block out sands and shales.

2 Calculate the SGR.

3 Determine the SGR threshold for seal/leak.

The first step in quantitative fault seal analysis is determining the detailed sand–shalecontent of the stratigraphy. A standardized method of sand–shale discrimination is anabsolute necessity. Digital logs and log analysis software simplify the process. Routinemethods of sand–shale discrimination are a basic part of analysis and most log analysissoftware (Doveton, 1986).

Step 1: Analyze logs

Page 27: Chap10

Evaluating Fault Seal • 10-27

The portion of the fault analyzed is adjacent to the Rd sand in the hanging wall. The SGRis the ratio of the total sand divided by the total shale that has moved past the fault. Thestratigraphy that has moved past this portion of the fault is marked in red. As the throwincreases, more of the stratigraphy below the Ru and Rd sands moves past the fault andthe SGR changes.

A plot of SGR vs. throw, taken from the preceding example, is shown in the following fig-ure. At small throws, very little shale has moved past the fault, the R sand is not com-pletely offset, and the SGR is high. As the throw increases, progressively more shale ismoved past the fault and the SGR decreases. At larger throws, a thick sand packagebegins to move past the fault and the SGR increases.

Quantitative Fault Seal Analysis, continued

Figure 10–23.

The following figure is an example of an SGR calculation for a portion of a fault at pro-gressive throws of 50, 100, and 200 ft. Sands are patterned; shales are black. The blacktrace is an SP log.

Step 2:Calculate theSGR

Figure 10–22.

Page 28: Chap10

10-28 • Evaluating Top and Fault Seal

Quantitative Fault Seal Analysis, continued

An SGR–throw plot shows the variation in SGR with increasing displacement for only onesmall area along the fault plane. In practice, SGR is calculated and mapped over the sur-face of the entire fault plane. The following figure shows the variation in SGR for eachpoint along the fault.

Step 2:Calculate theSGR(continued)

Figure 10–24.

Page 29: Chap10

Evaluating Fault Seal • 10-29

Quantitative Fault Seal Analysis, continued

The SGR threshold for a basin is defined by determining the SGR for known sealing andleaking faults in producing fields. The empirical threshold can then be used to assessprospects and plays and well as develop fields. SGR thresholds vary among basins.

The following example of an SGR threshold plot for the U.S. Gulf Coast directly assignsfault seal risk based upon the variation in SGR and seal behavior observed in a data setof approximately 160 faults. The plot shows the percent of faults cross sealing for oil as afunction of changing SGR. The actual SGR values have been deleted intentionally.

Above a specific SGR value, 100% of all faults in the Gulf Coast cross-leak oil. Below aspecific SGR value, 100% of the faults cross-seal oil. A narrow transition zone defines thethreshold for seal/leak. Separate thresholds exist for cross-seal/-leak, dip-seal/-leak, oil,and gas.

Determine SGRthreshold

Figure 10–25.

Page 30: Chap10

10-30 • Evaluating Top and Fault Seal

The limitations of quantitative fault seal analysis follow:• It applies only to faulted sand/shale sequences. It is not applicable to massive carbon-

ate, chert, or sand reservoirs. It has yet to be tested in interbedded shale–carbonatesequences.

• The seal/leak threshold SGR must be empirically calibrated for each basin, usingknown sealing and leaking faults. An SGR threshold for the Gulf Coast cannot be usedfor assessing prospects in the Gippsland basin. The confidence with which a seal can berisked is thus much greater in a production setting or mature basin than it is in a fron-tier setting.

• It does not apply to all structural styles, and specifically it does not necessarily apply tofaults in foreland fold and thrust belts or strike-slip basins. All basins in which quanti-tative fault seal analysis has been proven to date are dominated by detached or base-ment-involved normal faults.

• Cataclasis, diagenetic effects, localized fracturing, sharp changes in the permeability ordisplacement pressure of sands, reactivation of earlier normal faults in compression,"shale-outs," and the lack of lateral sand continuity in fluvial sequences can affect sealbehavior.

• The ability to predict fault seal behavior is only as good as the ability to predict thestratigraphy and structure. As with most variables in prospect assessment, uncertain-ties in structure and stratigraphy lead to a minimum, maximum, and most likely faultseal risk.

Quantitative fault seal analysis is a proven tool in numerous basins. There are, however,limitations.

Introduction

Limitations of Quantitative Fault Seal Analysis

Limitations

Page 31: Chap10

Evaluating Fault Seal • 10-31

Routine fault seal analysis integrates both fault plane profiles and quantitative fault sealanalysis. This discussion contains an example from a field in the Gulf Coast and demon-strates the application of routine fault seal analysis.

Introduction

Example of Routine Fault Seal Analysis: Gulf Coast

Procedure

The following figureis an FPP of a portionof a fault in a GulfCoast field. It showsthe juxtaposition ofsands along the faultand the observed dis-tribution of hydrocar-bons. Sands areshown in gray (hang-ing wall) and orange(footwall).

Construct anFPP

The table below outlines a procedure for analyzing fault seal behavior.

Step Action

1 Construct a fault plane profile (FPP).

2 Determine seal behavior of the fault.

3 Calculate the smear–gouge ratio.

4 Construct a smear–gouge ratio map.

5 Determine the smear–gouge ratio threshold.

Figure 10–26.

Page 32: Chap10

10-32 • Evaluating Top and Fault Seal

Examples of cross-leaking and cross-sealing portions of the fault are shown in the follow-ing figure, which is a detail of Figure 10–26. Juxtapositions of the 7100/7100 sands andthe 7200/7200 sands are cross leaking (left). These sand/sand juxtapositions have com-mon hydrocarbon contacts across the fault. The portion of the fault juxtaposing the7100/7200 sands, however, is cross sealing (right). Hydrocarbons in the 7200 sand in thehanging wall (Ru) are juxtaposed with water-wet sands of the 7100 sand in the footwall(Rd).

Determine sealbehavior

Example of Routine Fault Seal Analysis: Gulf Coast, continued

Figure 10–27.

Page 33: Chap10

Evaluating Fault Seal • 10-33

SGR is calculated over the surface of the entire fault. The following figure is a map of theSGR contoured on the FPP of the fault in Figure 10–26. SGR—and seal potential—varyover the surface of the fault as the throw and stratigraphy change. Areas of high SGR areshown in red and orange. Areas of low SGR are shown in purple and blue.

Calculate SGRand constructSGR map

Example of Routine Fault Seal Analysis: Gulf Coast, continued

Figure 10–28.

Page 34: Chap10

10-34 • Evaluating Top and Fault Seal

The SGR is calculated by “rubber-banding” the stratigraphy in adjacent well logs to fit thestratigraphic cutoffs on the FPP. The detailed log stratigraphy is necessary for reliablycalculating SGR. The rubber-banding corrects for apparent thicknesses in both the welland fault plane profile.

Calculate SGRand constructSGR map(continued)

Example of Routine Fault Seal Analysis: Gulf Coast, continued

Figure 10-29.

The SGR is calculated for each known sealing and leaking segment of the fault. The SGRsof two cross-leaking and cross-sealing segments of the fault are shown in the following fig-ure (location is on Figure 10–33). SGR values along the cross-leaking sand/sand juxtaposi-tions are high (red). Cross-leakage (right) occurs at both the 7100/7100 and 7200/7200juxtapositions. Cross-seal (left) occurs where the 7100 sand is juxtaposed with the 7200sand. This cross-sealing segment is characterized by low SGRs (blue). Cross-leakageoccurs where the SGR is high; cross-seal occurs where it is low.

Example:correlatingSGRs, sealbehavior

This same iterative process is used to construct SGR threshold plots and to calibrate thecritical SGR necessary for seal and leak. The SGR threshold plot of the Gulf Coast notedearlier in this subsection was constructed from a similar analysis of 160 known sealingand leaking fault segments. The constantly expanding database now includes 246 faultsin the Gulf Coast. The result is the ability to risk fault seal behavior with a high degree ofconfidence. Fewer than 10% of the faults analyzed in the Gulf Coast are exceptions to thispattern.

Calibrating SGRthreshold

Page 35: Chap10

Evaluating Fault Seal • 10-35

The sealing behavior of faults controls not only the entrapment of hydrocarbons but alsothe migration pathways into a trap. Fault seals control not only whether a trap retainshydrocarbons but also the volume of hydrocarbons available to migrate into a trap.

Migration pathway maps trace hydrocarbon movement between source and trap within acomplexly faulted field and between fields. Migration maps that consist solely of diparrows drawn on maps of the top reservoir can be very misleading. A migration pathwaymap must use the detailed information available from routine fault seal analysis.

Introduction

Subsection A4

Fault Seal and Migration Pathways

This subsection contains the following topics.

Topic Page

How Faults Control Trap Fill and Migration Pathways 10–36

How to Construct Migration Pathway Maps 10–38

Effect of Seal on Hydrocarbon Yield Estimates 10–39

In thissubsection

Page 36: Chap10

10-36 • Evaluating Top and Fault Seal

A sealing fault can trap hydrocarbons but also will act as a barrier for hydrocarbon migra-tion into traps beyond that fault. In addition, faults can act as baffles by deflecting hydro-carbons along migration pathways that may not be perpendicular to structural contours.

Introduction

How Faults Control Trap Fill and Migration Pathways

Dry traps, or traps with limited fill, may exist because hydrocarbons have been trappeddowndip along sealing faults. A fault can act as a barrier to (1) all hydrocarbons or (2)some of the hydrocarbons, allowing a limited volume to migrate. It can also act as a barri-er along part of the fault plane and as a conduit along other parts of the fault plane.

Controlling trapfill

Where complex fault systems exist between a trap and a source kitchen or between twotraps, migration pathways are correspondingly complex. Traps in this setting can havewidely different migration/fill or charge risks, depending upon fault seal behavior.

Controllingmigration

Where the dip of carrier beds is not perpendicular to faults, even cross-leaking faults canact as barriers as long as the permeability of the carrier bed is higher than that of thefault. Hydrocarbons can then migrate parallel to a fault rather than across the fault, eventhough the fault cross-leaks. This baffle effect can direct hydrocarbons away from poten-tial traps as well as toward others. Migration pathway maps are critical to prospectassessment.

Migrationparallel tofaults

The Hudson field, North Sea, is an excellent example of how faults control migration path-ways and charge (Hardman and Booth, 1991). The map and cross section in the following figure show that the first well, 210/24a-1, was located on the crest of an obvious structuralhigh. This well encountered water-wet Brent Group sands. A second well on the flank,210/24a-2, encountered oil shows suggestive of a local stratigraphic trap. Thirteen years after the initial well, the 210/24a-3 well discovered the Hudson field: a fault-dependent trap.

Example:Hudson field

Figure 10–30. After Hardman and Booth, 1991; courtesy Geological Society of London.

Page 37: Chap10

Evaluating Fault Seal • 10-37

The sealing fault trapped hydrocarbons in a flank fault compartment and preventedhydrocarbons from migrating into the more obvious structural high to the west. Predrillfault seal analysis and a migration pathway map would have correctly identified thesealing fault and would have placed a much greater risk on the success of the first well. Inthis case, new seismic data identified the sealing fault; however, numerous examples existwhere the same error is made with high-quality seismic data. Hydrocarbons do not simplymigrate into the crest of structural highs.

Example: Hudson field(continued)

How Faults Control Trap Fill and Migration Pathways, continued

In the Don field, North Sea, sealing faults prevent hydrocarbons from migrating into faultcompartments on the crest of a large structural high (Hardman and Booth, 1991).Instead, hydrocarbons are trapped in several fault compartments on the flank of thestructure against cross-sealing faults that have sand/sand juxtapositions. Wells in threefault compartments (211/18-5, 10, and 16) in the crest of the structure are dry.Hydrocarbons have either been trapped downflank or have been deflected to the south-west by sealing faults.

Example: Donfield

Figure 10–31. After Hardman and Booth, 1991; courtesy Geological Society of London.

Page 38: Chap10

10-38 • Evaluating Top and Fault Seal

A migration pathway map is a fundamental tool for prospect assessment and field devel-opment. In simple migration pathway maps, arrows are drawn perpendicular to structur-al contours. However, migration pathway maps must also define those faults and thoseparts of faults that juxtapose sand/sand as well as those faults that cross-seal/cross-leakand dip-seal/dip-leak.

Introduction

How to Construct Migration Pathway Maps

The following table lists the requirements for constructing a migration pathway map.

Requirement Explanation

Structure–depth maps Defines buoyant vectors for hydrocarbon migration

Fault plane profile Defines sand/sand and sand/shale juxtaposition and potential juxtaposedlithology leak points (JLLPs)

Quantitative fault plane analysis Defines real JLLPs, areas of cross-seal or cross-leak, and areas of dip seal or dip leak

Requirementsfor mapconstruction

Figure 10–32.

The figure on the rightillustrates how completelya fault can control migra-tion pathways across afault. In this fault planeprofile of a fault in a GulfCoast field, SGR was cal-culated and contouredonly on areas of sand/sandjuxtaposition. Areas ofsand/shale juxtapositionare uncolored. No hydro-carbon migration willoccur across these uncol-ored regions or across thesand/shale juxtapositionswith low SGRs (blue-pur-ple). These high SGRareas are essentially holesfor fluid migration in thefault plane. It will onlyoccur across the areas ofsand/sand juxtapositionwith high SGRs (red-orange).

Example: Faultcontrollingmigration

Page 39: Chap10

Evaluating Fault Seal • 10-39

Hydrocarbon yield models estimate the volume of hydrocarbons that can be generatedfrom a given volume of source rock and thermal history. These calculations are commonlycalibrated by comparing the volume of hydrocarbons that should have been generated in adrainage area with the volume actually trapped. This "calibration factor" then is usedroutinely in yield estimates.

Calibrationfactor

Effect of Seal on Hydrocarbon Yield Estimates

The logical error in many of these yield calibration studies is assuming that the hydrocar-bon volume trapped is the total volume that migrated into the trap. If the trap contains afault-dependent leak point, however, the volume that migrated in may have been muchgreater than the volume now trapped. The additional hydrocarbons spilled across a fault-dependent leak point and migrated updip. If these fault-dependent leak points are notidentified and we assume the trap contains all the hydrocarbons generated in the sourcekitchen, then calibration factors and yield estimates can be extremely misleading. Inaddition, hydrocarbons may spill from adjacent traps rather than being directly derivedfrom source kitchens.

Potential errors

Page 40: Chap10

10-40 • Evaluating Top and Fault Seal

Routine fault seal analysis is an important part of field development and field unitization,yet it is routinely ignored at considerable cost (Knutson and Erga, 1991; Jev et al., 1993).Faults control the following:• Vertical and lateral distribution of hydrocarbons within fault compartments• Communication between fault compartments• Movement of hydrocarbons during production

Introduction

Subsection A5

How Faults Affect Field Development

This subsection contains the following topics:

Topic Page

Hydrocarbon Distribution 10–41

Reservoir Simulations and Field Unitization 10–43

Fault Seal Breakdown During Production 10–44

In thissubsection

Page 41: Chap10

Evaluating Fault Seal • 10-41

Despite numerous examples of this type, we often erroneously assign an equal chance ofsome specific percent fill when assessing a prospect’s reservoirs and fault compartments.We must understand the following:• All reservoirs in a series of stacked sands do not have the same chance of being filled

with hydrocarbons.• Each fault compartment of a single reservoir does not have the same chance of being

filled with hydrocarbons.• We can predict (predrill) which sands and fault compartments contain hydrocarbons

and which are dry.

Assessingpercent fill

Faults control both the vertical and lateral distribution of hydrocarbons within fields.Some compartments and sands are dry; others contain subeconomic accumulations.Risking percent fill using fault seal analysis helps avoid needless wells.

Introduction

Hydrocarbon Distribution

The following figure is an example of faults controlling the distribution of hydrocarbonswithin a Gulf Coast field. The field contains 33 separate reservoir sands. At each mappedreservoir level, the field is divided into three major compartments (A, B, D) by faults.Compartment D is almost always empty, while the other compartments commonly containhydrocarbons. In only 5 of the 33 reservoirs does Compartment D contain hydrocarbons.

Compartment D is commonly empty because a series of JLLPs along the western portionof Fault A allows hydrocarbons to leak into Compartment B and then across the easternportion of Fault A into the next highest reservoir. Only when Fault A is cross sealing doesCompartment D contain hydrocarbons. Cross seal is created by shale-prone gouge withlow SGRs or sand/shale juxtaposition.

Distributionexample

Figure 10–33.

Page 42: Chap10

10-42 • Evaluating Top and Fault Seal

Hydrocarbon Distribution, continued

Understanding communication of fluids across faults lets us develop fields more efficient-ly. Routine fault seal analysis during production avoids needless wells and helps us posi-tion necessary infill wells for producing residual accumulations. Cross-sealing faults com-partmentalize a reservoir. Isolated compartments created by cross-sealing faults requireindividual production wells. In contrast, cross-leaking faults allow some fault-boundedcompartments to be produced from adjacent compartments and additional wells may beunnecessary.

Communicationof fluids acrossfaults

Page 43: Chap10

Evaluating Fault Seal • 10-43

Fault plane profiles and quantitative fault seal analysis are required for realistic reser-voir simulations. Neither the seal behavior, transmissibility, permeability, nor areas ofsand/sand juxtaposition are constant over the entire fault surface.

Introduction

Reservoir Simulations and Field Unitization

Faults can “pond” hydrocarbons and affect sweep and waterflood efficiency. Routine faultseal analysis may be required for producing residual oil accumulations missed by assum-ing a laterally continuous reservoir.

Movement ofhydrocarbons

Faults control changing pressure gradients within a field. Hydrocarbons move in responseto these pressure gradients and not necessarily in response to structural dip. Gas in theBeryl field, for example, migrated downdip during production in response to changingpressure gradients, pressure compartments, and migration pathways controlled by seal-ing and leaking faults (Knutson and Erga, 1991; Skerlec, 1997b).

Fault control ofpressuregradients

Faults can also leak over geologic time spans but seal during production time spans. Evena low fault-zone permeability may allow hydrocarbons to leak given a time span of 106

m.y. High production rates, however, creating pressure changes over a span of 1–10 years,will cause low-permeability fault zones to act as barriers to hydrocarbon movement. Across-leaking fault may develop different hydrocarbon contacts and different pressuresduring production. SGR thresholds for seal behavior may have to be calibrated separatelyfor exploration and for reservoir simulations.

Fault sealbehavioralteration

Fault seal is important in field unitization. Ignoring fault seal and depending solely onreservoir parameters and estimated hydrocarbon contacts can lead to extremely unequaldivision of reserves. The sealing behavior of faults controls both the original distributionof hydrocarbons in a field as well as the volumes of hydrocarbons produced from individ-ual fault compartments.

Fault seal andfield unitization

Page 44: Chap10

10-44 • Evaluating Top and Fault Seal

Introduction

Fault Seal Breakdown During Production

Changes in pressure differentials across faults can change seal behavior. A cross-sealingfault is cross sealing to a finite column of hydrocarbons. As the buoyant pressure increas-es at the crest of the hydrocarbon column, the buoyant pressure will ultimately exceed thedisplacement pressure of the fault zone and the fault will cross-leak (Smith, 1966; Buckand Robertson, 1996).

Pressure depletion during production can create large pressure differentials across a faultand consequent leakage. This process occurs when the pressure depletion is sufficientlyrapid, or the fault has sufficiently low permeability, to cause a large pressure differentialacross the fault. The change from sealing to leaking is not caused by any mechanical rup-turing but simply by the relative change in buoyant pressure vs. displacement pressure.

Pressuredifferentialsand leakage

The following figure shows the pressure depletion curves for two wells separated by aninitially cross-sealing fault. The pressure depletion curve for well A is shown in light gray(top); that for well B, in black (bottom). The buoyant pressure at the crest of the oil col-umn against the fault seal in well A remains constant as the pressure in well B decreases.The pressure differential (∆P) increases until the displacement pressure of the fault zoneis exceeded and the fault begins to cross-leak. A fault in the Beryl field has broken downduring production (Buck and Robertson, 1996). A fault in the Akaso field, Nigeria, mayhave undergone this type of breakdown with a differential pressure of 4137 kPa (600 psi)(Jev et al., 1993).

Example

Figure 10–34.

Page 45: Chap10

Evaluating Top Seal Integrity • 10-45

Fracturing can destroy top seal integrity. This section discusses two practical techniquesfor evaluating loss of top seal integrity resulting from strain or hydraulic fracturing.Sufficiently intense deformation in excess of top seal ductility can fracture a top seal. Inaddition, sufficiently high pore pressures in excess of the fracture pressure can inducenatural hydraulic fracturing.

Introduction

Section B

Evaluating Top Seal Integrity

This section contains the following subsections.

Subsection Topic Page

B1 Strain Analysis of Top Seals 10–46

B2 Overpressure and Natural Hydraulic Fracturing 10–56

In this section

Page 46: Chap10

10-46 • Evaluating Top and Fault Seal

Loss of top seal integrity by high strain and fracturing is an important cause of dry holesand partially filled traps (Skerlec, 1982, 1992). Although fracturing is commonly predictedfrom stress, the advantage of strain analysis is that it provides a practical, predrill tool forevaluating prospects using seismic data.

Introduction

Subsection B1

Strain Analysis of Top Seals

Even a small amount of fracturing can result in staggering leakage rates. The volume ofhydrocarbon that could be lost from a typical North Sea field, assuming a fractured topseal with a relatively low fracture permeability of only 0.05 md, is more than 100 billionbbl/m.y. (Skerlec, 1990). Fracture permeability can be as high as tens of darcys (Stearnsand Friedman, 1972) and leakage rates consequently can be much higher.

Natural seeps confirm these high rates of hydrocarbon loss. The Palos Verde fault inSanta Monica Bay, California, seeps oil at the rate of 10–15 bbl/d or more than 5 billionbbl/m.y. The Coal Point seep in the Santa Barbara Channel, California, is leaking 50–70bbl/d or more than 25 billion bbl/m.y. (Wilkinson, 1971).

Rates ofleakage

This subsection discusses techniques for evaluating the loss of seal integrity resultingfrom deformation and contains the following topics.

Topic Page

Seal Ductility 10–47

Estimating Ductility of Top Seals 10–48

Estimating Strain in Top Seals 10–51

Example: Evaluating Top Seal Integrity 10–53

In thissubsection

Page 47: Chap10

Evaluating Top Seal Integrity • 10-47

The most important mechanical property for evaluating seal integrity is ductility. Ductilerocks make good top seals; brittle rocks make poor top seals. Shales and salt are two ofthe most ductile rock types and, not surprisingly, two of the most common top seals(Grunau, 1987).

Introduction

Seal Ductility

Ductility is the amount of strain a seal can withstand before brittle failure and the loss oftop seal integrity. Rocks with an extremely high ductility can deform without brittle fail-ure. On the other hand, rocks with low ductility can accommodate only a small amount ofstrain before fracturing. A seal can be brittle but unfractured; a seal can be ductile butfractured. Fracture depends upon whether the strain exceeds the seal ductility.

What isductility?

Seal ductility is controlled by at least nine different variables. The table below lists thesevariables and briefly notes how they control ductility.

Variable Control on Ductility

Lithology Grain mineralogy and cement type control ductility. Brittle seals include dolostone,quartzite, anhydrite, and some shales. Ductile seals include halite, some shales, and some limestones.

Composition Not all limestones or shales have the same ductility. Compositional variations suchas total organic carbon (Chong et al., 1980) and clay mineralogy (Corbett et al.,1987) change ductility.

Confining pressure Increasing confining pressure increases ductility.

Pore pressure Increasing pore pressure decreases ductility.

Fluid composition The presence or absence of fluids and their composition affects ductility.

Temperature Increasing temperature increases ductility.

Strain rate High strain rates decrease ductility.

Time Ductility changes with time as seals undergo burial and diagenesis.

Compaction state Ductility decreases with progressive compaction and diagenesis.

Variables thatcontrol ductility

Page 48: Chap10

10-48 • Evaluating Top and Fault Seal

The ductility of a top seal can be estimated (1) by using laboratory data and log-deriveddensity values that reflect the compaction state or (2) by comparing the strains in testedtraps associated with successes and failures.

Introduction

Estimating Ductility of Top Seals

The ductility of a shale top seal is a function of compaction state. Uncompacted, low-den-sity shales are extremely ductile and can thus accommodate large amounts of strain with-out undergoing brittle failure and loss of top seal integrity. Highly compacted, denseshales are extremely brittle and undergo brittle failure and loss of top seal integrity withvery small amounts of strain.

The following figure shows the relationship between ductility and density for 68 shales.The ductility of the shales was measured in the laboratory at confining pressures of 1,200, and 500 kg/cm2. All samples were deformed in compression.

Shale ductilityand density

Figure 10–35. Data from Hoshino et al., 1972.

Figure 10–35 illustrates the following:• Shales with a density less than approximately 2.1 g/cm3 deform ductilely and not by

brittle failure at the range of confining pressures found in most sedimentary basins.• Shales with a density greater than 2.1 g/cm3 will undergo brittle failure given sufficient

strain.• The ductility can be inferred from the density. Denser shales are more brittle and can

withstand less strain before fracturing. Less-dense shales are more ductile and canwithstand larger strains before fracturing.

• Ductility increases with increasing confining pressure.

Page 49: Chap10

Evaluating Top Seal Integrity • 10-49

Estimating Ductility of Top Seals, continued

The ductility of a shale top seal decreases with progressive burial, compaction, and diage-nesis within a sedimentary basin. The mechanical properties are not constant but changewith progressive burial as the top seal is converted from a mud to a rock.

The ductility of a shale top seal also increases in response to increasing confining pres-sure. Thus, a shale with constant mechanical properties will have a lower ductility atshallow depths than at greater depth. Since a shale top seal does not have constantmechanical properties with progressive burial, compaction decreases ductility at the sametime as confining pressure increases ductility.

Shale ductilityvs. depth

The following figure shows the change in density and ductility of shales with increasingdepth. Laboratory data are plotted on a normal shale compaction curve showing densityvs. depth. The figure shows the ductility of each shale at that depth or confining pressure,with ductile samples shown by gray circles and brittle samples shown by black circles.Ductile shales did not fracture; brittle shales fractured. A low-density shale at a depth of500 m is more ductile than a highly compacted shale at a depth of 5000 m in the center ofthe basin. Identical traps, one in the graben deep and one on an adjacent marginal plat-form, have different seal risk.

Shale density–ductility vs.depth

Figure 10–36. Data from Hoshino et al., 1972.

This technique has been used on other seal lithologies besides shales, including siltstones,marls, and chalks (Skerlec, 1990). There is no relationship between the density and theductility of sandstones where high permeability and diagenetic effects alter the more sim-ple relationship observed in less permeable rocks.

Use in otherlithologies

Page 50: Chap10

10-50 • Evaluating Top and Fault Seal

Ductility changes not only with depth of burial but also with time and progressive subsi-dence. A shale top seal now buried at 4000 m and having a density of 2.6 g/cm3 was onceburied at a more shallow depth and had a lower density. This now-brittle seal was onceductile.

Ductility vs. time

Estimating Ductility of Top Seals, continued

To predict paleoductility, we must know both the density and the confining pressure atthe time of deformation. A database of top seal mechanical properties over a range of per-tinent confining pressures is a basic tool for seal analysis.

Predictingpaleoductility

Ductility–time plots can be constructed from shale compaction curves and burial historycurves. Burial history curves give the depth of burial of a top seal at a specific time. Shalecompaction curves let us infer the shale density at a specific depth of burial and time.

The following figure is a ductility–time plot for an Upper Jurassic top seal in the CentralGraben, North Sea. The plot shows the paleodensity and inferred paleoductility duringprogressive burial of shales at the 141- and 151-m.y. sequence boundaries. Prior toapproximately 100 m.y., the Late Jurassic shale top seal had a density of < 2.1 g/cm3 andwas ductile. Strain prior to 100 m.y. would not contribute to seal risk. Any deformationoccurring after 100 m.y. could have caused fracturing, given sufficiently high strains.

Ductility–timeplots

Figure 10–37.

Page 51: Chap10

Evaluating Top Seal Integrity • 10-51

Once the ductility or paleoductility of a seal is known, the amount of strain that hasaffected the seal must be determined. One of the most useful techniques is incrementalstrain analysis. Originally applied to fractured reservoirs (Watts, 1983), the technique isequally applicable to top seal analysis (Skerlec, 1982, 1992; Koch et al., 1992).

Introduction

Estimating Strain in Top Seals

Incremental strain analysis measures the change in line length of sequence boundarieson depth-converted seismic lines to calculate strain magnitude as well as the strain occur-ring at different time intervals. The advantages of this technique are that it can beapplied to prospects before drilling and it yields a quantitative estimate of strain that canbe compared directly with strain measurements of top seals in the laboratory.Incremental strain analysis yields not only a quantitative strain value but also showshow strain has varied with time.

Incrementalstrain analysis

The following figure is a simple cross section of a low-relief salt structure showing threeseismic sequences. The figure shows how strain is calculated: by comparing the originalundeformed line length with the deformed line length between two arbitrary points, A and B. In this figure, l0 is the initial, undeformed line length; l1 is the deformed linelength of the 60-m.y. sequence boundary; and l2 is the deformed line length of the 131-m.y. sequence boundary. The strain (ε) on the 131-m.y. sequence boundary is ε = (l2 – l0)/l0

= 2.5%. The strain on the 60-m.y. sequence boundary is ε = (l1 – l0)/l0 = 0.5%. All strainsare extensional.

Calculatingincrementalstrain

Figure 10–38.

Page 52: Chap10

10-52 • Evaluating Top and Fault Seal

This method provides both an estimate of quantitative strain and the timing of strain.The incremental strain is the strain occurring during the time interval between twosequence boundaries. In the preceding figure, 0.5% strain occurred between 60 Ma andthe present, while 2.5% strain occurred between 131 Ma and the present. Consequently,2.0% strain occurred between 60 and 131 Ma. Strain–time plots display the strain affect-ing the top seal from the onset of deformation to the present.

Calculatingincrementalstrain(continued)

Estimating Strain in Top Seals, continued

Analysis requires decompaction of the sediments to correct for the apparent strain causedby compaction. In the Central Graben, the effect is generally minimal.

Correction forcompaction

How much strain can a top seal withstand? In a mature basin, risk assessment of thestrain threshold at which failure occurs is determined empirically by analyzing successesand failures. The empirical threshold varies from basin to basin. In a frontier setting, esti-mates of the strain required for failure must be based on (1) laboratory data on seal prop-erties and (2) estimates of the confining pressure and seal ductility at the time of trap for-mation.

Empiricalthreshold fortop seal failure

Incremental strain analysis provides only an estimate of strain. More detailed calcula-tions of stain may be warranted. Incremental strain analysis is an average of strain overthe entire structure. Local areas of high strain can have important control on spill points.Recent analyses have taken shorter line-length increments over the structure to calculatesmall-scale variations in strain (Koch et al., 1992). In addition, the method assumes fixedend points, whereas a limited amount of flexural slip occurs.

Limitations ofincrementalstrain analysis

Page 53: Chap10

Evaluating Top Seal Integrity • 10-53

Following is an example of evaluating, or risking, top seal integrity using two traps in theCentral Graben, North Sea. One trap was dry; one trap is now a producing field. Both arelow-relief salt structures, a common trap style in the Central Graben, with UpperJurassic reservoirs sealed by Upper Jurassic–Lower Cretaceous shales. A seismic lineacross both traps is shown in the following figure. Trap A is at the left; trap B is at thecenter.

Introduction

Example: Evaluating Top Seal Integrity

Figure 10–39. Courtesy Esso Exploration and Production U.K.

Incremental strain was calculated from depth-converted seismic profiles over the mosthighly strained portion of each structure using a series of seismic sequence boundariesthat could be mapped throughout the basin. The sea floor (0 Ma sequence boundary) rep-resents the original, undeformed line length across each structure.

Method

Page 54: Chap10

10-54 • Evaluating Top and Fault Seal

The following strain–time plots show the incremental strains affecting the top seal aboveeach trap and the wide variation in strain magnitude as well as strain history in thesetraps. Trap A has a maximum strain of 4.5%, which contributes to the risk of top sealfracturing. The time of maximum strain occurred 97–60 Ma. Today, trap A is a dry holebecause of fracturing and loss of top seal integrity. Trap B has a maximum strain of only1%, which occurred before 100 Ma, and only a 0.2% strain that occurred 97–60 Ma. Today,trap B is a producing structure with an intact top seal capable of trapping hydrocarbonsin the underlying Upper Jurassic sands.

Strain–timeplots

Example: Evaluating Top Seal Integrity, continued

Figure 10–40.

Paleoductility analysis demonstrates that the top seals in these structures were extreme-ly ductile prior to 100 Ma and would not have fractured within the range of strains devel-oped. Since 100 Ma, these same seals were sufficiently compacted so that fracturing couldhave occurred if the strain exceeded the ductility of the top seal. The only strain that con-tributes to seal risk is the strain that occurred after 100 Ma.

Paleoductilityanalysis

Page 55: Chap10

Evaluating Top Seal Integrity • 10-55

Incremental strain analysis of a group of successes and failures in the Central Grabendefines a strain threshold at approximately 1.6%. In the following figure, traps with amaximum post-100 Ma incremental strain > 1.6% are dry (traps H, I, J). Traps with astrain < 1.6% are still intact and contain hydrocarbons (traps B–G). Trap A, a structurewith very low strain, is dry because of natural hydraulic fracturing. Traps that were drydue to other causes such as lack of reservoir or source are not included in this plot.

Strain thresholdfor failure

Example: Evaluating Top Seal Integrity, continued

Figure 10–41.

This empirical threshold provides an important tool for evaluating top seals. Prospectscan be evaluated before drilling using incremental strain analysis. In this case, high-riskprospects are those with strains exceeding the threshold of 1.6%. Low-risk prospects arethose with strains below the threshold.

Routine top seal analysis changes the ranking of prospects as well as the explorationstrategy in the Central Graben. Large prospects, normally at the top of the ranking,unfortunately may also be large-strain, high-risk structures. The chances of success canbe increased by drilling low-strain, low-risk structures. Note that prospect size and strainmagnitude are not necessarily related. Large prospects could be low-relief, low-strainstructures. Continued drilling in the Central Graben has confirmed this strain thresholdwith a larger sample of tested traps (Koch et al., 1992).

Importance toexploration

Incremental strain analysis is not limited to analyzing low-relief salt structures like thoseused in the above example. It can be applied to any structural style involving deformationof the top seal, including traps in foreland fold-and-thrust belts, Gulf Coast growth faults,and North Sea normal faults. Deformation may involve shortening or extension.

Applying strainanalysis tostructural styles

Page 56: Chap10

10-56 • Evaluating Top and Fault Seal

Top seals can be hydraulically fractured by high pore fluid pressure. This section discuss-es evaluating natural hydraulic fracturing and presents pertinent examples.

Introduction

Subsection B2

Overpressure and Natural Hydraulic Fracturing

This subsection contains the following topics.

Topic Page

Natural Hydraulic Fracturing of Top Seals 10–57

Fracture Threshold in the Real World 10–59

Natural Hydraulic Fracturing Example, North Sea 10–61

Overpressure and Hydrocarbon Distribution, Gulf Coast 10–62

In thissubsection

Page 57: Chap10

Evaluating Top Seal Integrity • 10-57

Fracturing and consequent loss of top seal integrity can occur by increasing pore pres-sure. High pore pressure can overcome the normal stresses that keep fractures closed.Similar fracturing is artificially induced during leak-off tests, well stimulations, and sub-surface waste disposal (Evans, 1996).

Introduction

Natural Hydraulic Fracturing of Top Seals

High pore pressure has fractured the top seal and lost hydrocarbons in several basins,including the North Sea (Skerlec, 1982, 1992; Caillet, 1993; Leith et al., 1993), theNorwegian Sea (Ungerer et al., 1990), and the Malay basin (Scharr, 1976). The process isundoubtedly more widespread. Loss of top seal integrity due to natural hydraulic fractur-ing also appears to control the risk economics and vertical distribution of hydrocarbons inthe Gulf Coast (Fertl and Leach, 1988; Leach, 1993a,b).

Importance ofhydraulicfracturing

The overpressure required to cause fracturing is traditionally calculated by determiningthe theoretical fracture pressure, Pf (Hubbert and Willis, 1957):

Pf = α σ3 + p

where:

Pf = theoretical fracture pressureσ3 = effective least principal stress or confining pressurep = pore pressureα = poroelastic constant, assumed to be 1 in most analyses (see Engelder and

Lacazette, 1990)

Fracture pressure is the fluid pressure necessary to overcome the normal stress thatkeeps the fractures closed.

Theoreticalfracturepressure, Pf

Page 58: Chap10

10-58 • Evaluating Top and Fault Seal

Natural Hydraulic Fracturing of Top Seals, continued

Use the steps outlined in the following table to calculate Pf.

Step Action Method

1 Use density logs to calculate overburden stress.For example, a 1-cm cube with a density of 2.4g/cm3 exerts an overburden stress of 2.4 g/cm2 atthe base of the cube.

2 Calculate the ratio from leak-off tests. Take caresince leak-off tests may report the pressure valueeither prior to or after the fracture pressure point(Eaton, 1969). Leak-off tests are also commonlytaken where casing has been set and may reflectthe mechanical properties of the cement casingrather than the wall rock. Alternatively, Poisson’sratio can be estimated from available laboratorydata (Lama and Vutukuri, 1978). Poisson’s ratioincreases with depth to approach a maximum of0.5.

3 Pore pressure can be determined from measure-ments or regional pressure maps or estimatedfrom burial history (Mann and Mackenzie, 1990).It may be necessary to predict paleopore pressure.

4Solve the equation σ3 = (σ1 – p )

5 Solve the equation Pf = α σ3 + p. The fracturepressure is commonly expressed as a gradient,and the equation becomes Pf/Z = α σ3/Z + p/Z,where Z is depth.

Calculating Pf

Calculate σ1, over-burden stress.

Determine ν,Poisson’s ratio.

Determine p, porepressure.

Calculate σ3, effective confiningpressure.

Calculate Pf, theoretical fracturepressure.

Variations on this equation as well as empirical relationships are common (Hubbert andWillis, 1957; Matthews and Kelly, 1967; Eaton, 1969; Breckles and Van Eekelen, 1982;Brennan and Annis, 1984). An alternative method of determining the principal stressesand fracture gradient is through the use of borehole deformation (Bell, 1990; Evans andBrereton, 1990).

Other ways tocalculate Pf

ν(1 – ν)

Page 59: Chap10

Evaluating Top Seal Integrity • 10-59

We don’t really know how high the pore pressure must be to induce fracturing. In fieldexamples, failure seems to occur below the fracture pressure (Skerlec, 1982, 1992; Dutta,1987; Lerche, 1990; Ungerer et al, 1990). Risk increases as overpressure increases rela-tive to Pf, but failure may occur below the theoretical fracture pressure. Other factors, notyet understood, control the point at which failure occurs.

Theoretically, fracturing occurs when the pore pressure reaches Pf. However, Pf increasesas pore pressure increases. Although the theory generally is described in terms of the porepressure needed to overcome the horizontal stress keeping the fractures closed, in prac-tice the pore pressure must approach the lithostatic pressure for brittle failure to occur(Lorenz et al., 1991).

The following figure charts Pf vs. pore pressure for a range of overburden stress gradients(0.5–1.0). The pore pressure equals or exceeds Pf only when the pore pressure is equal toor greater than the lithostatic stress. Fracture occurs when Pf equals the pore pressure.

Introduction

Fracture Threshold in the Real World

Figure 10–42.

Although the variables used to calculate the fracture pressure commonly are generalizedas a smooth curve, there are local variations and departures from this generalized curve.The principal horizontal stress, σ3* and Poisson’s ratio depend on lithology and bed geom-etry as well as depth and compaction (Lorenz et al., 1991).

Stress andPoisson’s ratiovariation

Pore pressure (p) is not always a smooth curve (Engelder and Leftwich, 1993).Decompaction is controlled in part by the permeability of sediment layers. Shales adja-cent to high-permeability layers may undergo rapid decompaction while thick shalesequences with no immediate access to high-permeability layers will decompact moreslowly and have high pore pressures. Local variations in pore pressure may be importantin evolving seal integrity during basin subsidence.

Pore pressurevariations

Page 60: Chap10

10-60 • Evaluating Top and Fault Seal

Pore pressure alone does not control hydraulic fracturing. Changes in the overburdenstress change the theoretical fracture pressure and seal risk. For example, water depthalters the overburden stress and therefore Pf. The figure below compares the fracture gra-dient pressure for the case of a well on land and the same well with an additional 298 ft(100 m) of water column. The water column substitutes low-density water for high-densi-ty rock. The result is a shift of Pf to lower values. If the water depth were sufficientlygreat, there would be an increased likelihood of hydraulic fracturing with no change inpore pressure.

Similarly, facies changes within a basin can alter the density distribution in the sedimentcolumn and seal risk. A facies change from dense carbonates to less dense siliciclasticschanges the overburden stress gradient. A higher pore pressure is required to fracture atop seal in the denser sediment column. Seal risk is greater in the less dense sedimentcolumn. The overburden stress gradient and seal risk similarly change with progressivesubsidence and compaction.

Effect of waterdepth,stratigraphy, and facieschanges

Fracture Threshold in the Real World, continued

Figure 10–43.

Page 61: Chap10

Evaluating Top Seal Integrity • 10-61

A trap in the Central Graben, North Sea, appears to have been bled dry by naturalhydraulic fracturing. This trap is a low-relief salt structure with a dry hole and has less-than-sufficient strain to fracture the top seal.

Trap description

Natural Hydraulic Fracturing Example, North Sea

Figure 10–44. Courtesy Esso Exploration and Production, U.K.

Evidence for hydraulic fracturing and seal rupture include the following:• A fossil oil-–water contact exists at the synclinal spill point of the trap. The trap was

once filled with hydrocarbon, but something happened after trap fill to bleed off hydro-carbons.

• Hydrocarbon shows exist throughout the sediment column above the reservoir. Thesehydrocarbons are direct evidence of a breached top seal.

• The pore pressure is close to the theoretical fracture pressure. In adjacent fields, thepore pressure is significantly less than Pf.

• Cores show vertical, open extension fractures rather than the more common shear frac-tures found in adjacent fields. These fractures are identical to fractures produced byhydraulic fracturing.

Evidence

Figure 10–44 shows that thepore pressure is close to thefracture pressure at the baseof the top seal (approximately1300 ft). The pore pressure isinferred from mud weightand RFT measurements(solid dots). Leak-off tests(LOT) help constrain the frac-ture pressure. The fracturepressure is close to the litho-static pressure or overburdenpressure.

Page 62: Chap10

10-62 • Evaluating Top and Fault Seal

Studies of the hydrocarbon distribution in the Gulf Coast relative to the top of overpres-sure suggest hydraulic fracturing and seal integrity influence the vertical distribution ofhydrocarbons, success ratios, and seal risk (Fertl and Leach, 1988; Leach, 1993a,b).

Introduction

Overpressure and Hydrocarbon Distribution, Gulf Coast

The chance of finding an economically successful accumulation in the Gulf Coast decreas-es with depth. This decrease is a result of several changes, including reservoir quality andmigration pathways. Most importantly, however, the distribution of hydrocarbons is close-ly related to the top of the overpressure zone. The figure shows the distribution of oil andgas production from more than 20,000 wells. Most hydrocarbons are found near or slight-ly above the top of the overpressured zone. The chance of success is reduced by 90–95% atdepths of 2,000–5,000 ft below the top of overpressure (Leach, 1993b).

Success ratevs. depth

Figure 10–45. After Leach, 1993b; courtesy Oil & Gas Journal.

Page 63: Chap10

Evaluating Top Seal Integrity • 10-63

The distribution of hydrocarbons relative to the top of overpressure changes as reservoirdepth changes is shown in the figure below. This pattern is consistent with loss of top sealintegrity because of hydrofracturing rather than a simple loss of reservoir quality or othervariable. Deeper reservoirs have the peak gas occurrence below the top of overpressure.This is consistent with the changing Pf with depth. As depth increases, confining pressureincreases and thus the amount of pressure required to fracture the top seal (Pf) increases.As fracture pressure increases, the depth of the first intact top seal increases.

The relationship between depth of reservoir and depth of maximum cumulative produc-tion below the top of overpressure is fundamental to seal risk. Further work is required torelate the distribution of hydrocarbons to actual fracture pressure rather than simplydepth below top of overpressure.

Hydrocarbondistribution vs.overpressure

Overpressure and Hydrocarbon Distribution, Gulf Coast, continued

Figure 10–46. After Leach, 1993b; courtesy Oil & Gas Journal.

Page 64: Chap10

10-64 • Evaluating Top and Fault Seal

An intact top seal is a seal that has not undergone fracturing. An intact top seal can traponly a finite column of hydrocarbon. In theory, the height of hydrocarbon column that canbe trapped can be calculated from the capillary properties of the seal pores system andthe physical character of the hydrocarbons and pore fluids. This venerable approach toanalyzing top seals has been covered in depth by Berg, 1975; Schowalter, 1979; Watts,1987; and Vavra et al., 1992.

Here, we will discuss briefly the three steps involved in evaluating seal capacity of intacttop seals:

Step Action

1 Determine fluid properties, including density, interfacial tension, andcontact angle, at reservoir conditions.

2 Determine displacement pressure of the seal and reservoir.

3 Calculate seal capacity.

Introduction

Section C

Evaluating Intact Top Seal

The steps in the procedure above as well as problems that can occur are discussed in thefollowing subsections.

Subsection Topic Page

C1 How Capillary Properties Control Seal 10–65

C2 Estimating Displacement Pressure 10–69

C3 Seal Capacity 10–75

C4 Pitfalls and Limitations of Estimating Seal Capacity 10–82

In this section

Page 65: Chap10

Evaluating Intact Top Seal • 10-65

Hydrocarbons invading the pore space of a seal must displace the pore fluids. The pres-sure necessary to force the hydrocarbons into the seal and form a continuous filament isthe displacement pressure. The pressure which forces the hydrocarbons into the seal isthe buoyant pressure of the hydrocarbon phase. The trapping capacity of a top seal isthe balance between the displacement pressure and the buoyant pressure. When thebuoyant pressure exerted by the hydrocarbon column exceeds the displacement pressureof the seal, the seal leaks. For example, a shale top seal that could seal a 100-m column ofoil might leak if the column increased to 101 m.

Introduction

Subsection C1

How Capillary Properties Control Seal

This subsection contains the following topics.

Topic Page

Buoyant Pressure 10–66

Top Seal Displacement Pressure 10–67

Calculating Maximum Hydrocarbon Column 10–68

In thissubsection

Page 66: Chap10

10-66 • Evaluating Top and Fault Seal

The buoyant pressure (Pb) exerted by a column of hydrocarbons against the overlying sealis as follows:

Pb = (ρw – ρh) g h

Or, in the mixed units of the oil field:

Pb = (ρw – ρh) 0.433 h

where:

ρw = density of water, g/cm3

ρh = density of hydrocarbon, g/cm3

g = acceleration due to gravity, cm/sec2

h = thickness of the hydrocarbon column, ft

Calculation

Buoyant Pressure

The following figure is a pressure–depth plot through an oil column trapped beneath a topseal. The buoyant pressure of the oil column at any depth is the difference in pressurebetween the oil gradient and the water gradient. The maximum buoyant pressure (Pmax)is at the crest of the oil column. The buoyant pressure is zero at the free water level(FWL).

Pressure–depthplot

Figure 10–47.

Page 67: Chap10

Evaluating Intact Top Seal • 10-67

Displacement pressure (Pd) is the pressure necessary to force hydrocarbons into the porespace of a rock and form a continuous hydrocarbon filament (Schowalter, 1979).Displacement pressure, which is measured in dynes/cm2, can be calculated by the follow-ing formula:

Pd = (2γ cos θ)R

where:

γ = interfacial tension, dynes/cm θ = contact angle or wettability, degrees R = pore throat radius, cm

Calculation

Top Seal Displacement Pressure

The displacement pressure of a seal depends on both the physical character of the seal(pore throat radius and pore throat size distribution) and the physical character of thehydrocarbons (interfacial tension and wettability).

Variables

The wettability, or contact angle θ, is 0° for hydrocarbon/water (Berg, 1975). If the wettingphase is oil or gas rather than water, the contact angle can range from 0 to 180°. Not allrocks are water wet, and oil-wet seals may not be as rare as commonly thought (Cuiec,1987). Organic-rich sediments may be source, seal, and oil wet.

Wettability

Hydrocarbon/brine interfacial tension values typically range from 15–72 dynes/cm (Vavraet al., 1992; Schowalter, 1979; Watts, 1987). Laboratory studies provide interfacial tensionvalues for a range of gas and oil compositions (Firoozabadi and Ramey, 1988).

Interfacialtension

Page 68: Chap10

10-68 • Evaluating Top and Fault Seal

The maximum column of hydrocarbons that can accumulate under a seal occurs when Pb

equals Pd. If Pb exceeds Pd, the hydrocarbons will leak through the seal. If Pb is less thanPd, it would be possible to seal a greater column of hydrocarbons.

Therefore, for the maximum hydrocarbon column

Pb = Pd

(ρw – ρh) g h = 2γ cos θR

h =(2γ cos θ)R(ρw – ρh) g

or, in mixed units,

h = Pd(ρw – ρh) 0.433

where:

h = height of the hydrocarbon (ft)

Calculation

Calculating Maximum Hydrocarbon Column

The height of the hydrocarbon column (h) in the above equations is a theoretical maxi-mum. The actual height is less because of the effect of the reservoir. If the reservoir itselfhad zero displacement pressure, the height of the hydrocarbon column would attain thetheoretical maximum and the OWC would coincide with the free water level. If the reser-voir has some displacement pressure greater than zero, then the height of the hydrocar-bon column is less than the theoretical maximum, or

h = (Pds – Pdr )ρw – ρh0.433

where:

Pds = displacement pressure of the sealPdr = displacement pressure of the reservoir

Effect on sealcapacity

Page 69: Chap10

Evaluating Intact Top Seal • 10-69

In practice, the displacement pressure is estimated from laboratory measurements. Inreal rocks, the single pore-throat radius of the equations on page 10–68 is replaced by acomplex pore-throat size distribution.

Introduction

Subsection C2

Estimating Displacement Pressure

The following table lists techniques commonly used to estimate the displacement pres-sure of a top seal.

Technique Needed for the technique

Mercury injection measurements Cores or cuttings

Log analysis Logs and database of laboratory measurements

Sedimentary facies Facies maps and database of laboratory measurements

Pore-size distribution Thin sections, cores, or cuttings

Estimatingtechniques

This subsection contains the following topics which correlate with the techniques listed inthe table above.

Topic Page

Measuring Pd Using Mercury Injection 10–70

Estimating Seal Capacity from Sedimentary Facies and Well Logs 10–72

Estimating Pd from Pore Size 10–74

In thissubsection

Page 70: Chap10

10-70 • Evaluating Top and Fault Seal

The displacement pressure is routinely inferred by forcing mercury into the pore space ofa sample (cores or cuttings) and measuring the percent of mercury saturation vs. increas-ing pressure.

Introduction

Measuring Pd Using Mercury Injection

This figure shows a typicalmercury capillary curve for asandstone. Mercury is firstforced into the largest connect-ed pore throats. Saturationincreases with increasing pres-sure as mercury continues tobe forced into progressivelysmaller pore throats.

Procedure

Figure 10–48. After Schowalter, 1979; courtesy AAPG.

Displacement pressure (Pd) is defined as the pressure necessary to form a continuoushydrocarbon filament in the pore space of the seal. It is commonly inferred from the injec-tion pressure at 10% saturation (Schowalter, 1979) for two reasons: 1. Most reservoirs have a pronounced plateau along which saturation rapidly increases.

The pressure at 10% or 40% saturation gives a similar Pd. 2. Measured saturations required to create a continuous hydrocarbon filament range

from 5–17% with an average of 10% (Schowalter, 1979).

Alternatively, some workers define Pd as the pressure at the first inflection point of capil-lary curve (Katz and Thompson, 1986). Figure 10–49 above shows the inferred Pd at both10% saturation and at the inflection point.

Values of Pd

Page 71: Chap10

Evaluating Intact Top Seal • 10-71

Measuring Pd Using Mercury Injection, continued

Samples for mercury injection laboratory analysis can include cores as well as cuttings.Measurements made from cuttings do not yield the same value as those from cores, sothey require an empirical correction factor that ranges from 15–250 psi (Sneider andNeasham, 1993).

Seals with low permeability and small pore throats may require longer equilibrationtimes during mercury injection (Vavra et al., 1992).

Cores, cuttings,and low-permeabilityrocks

Since laboratory measurements of Pd are given in the air–mercury system rather than theoil–water or gas–water systems, we must convert from Pdm, using mercury, to Pdh, orhydrocarbons:

Pdh =γh cos θh Pdm

γmcos θm

Displacement pressures measured in the air–mercury system are then converted to thehydrocarbon–water system at subsurface conditions. To convert, we must know the tem-perature, pressure, wettability, and coefficient of interfacial tension for the hydrocarbonphase. These parameters are commonly inferred from the composition, gas–oil ratio, andAPI gravity (Schowalter, 1979; Vavra et al., 1992). For the air–mercury system, the wetta-bility of mercury is 140° (cos 140 = 0.766). The coefficient of interfacial tension for mer-cury is 485 dynes/cm (Vavra et al., 1992).

Convertinglaboratorymeasurements

Page 72: Chap10

10-72 • Evaluating Top and Fault Seal

Displacement pressure has been inferred from the relationship between capillary proper-ties and sedimentary facies (Vavra et al., 1992; Shea et al., 1993). Two examples demon-strate the ability to construct regional facies maps and then assign maximum limits tothe amount of hydrocarbon that could be trapped beneath these seals.

Introduction

Estimating Pd from Sedimentary Facies and Well Logs

The top seal and capillary properties of rocks from the Ardjuna basin (offshore Java,Indonesia) are related to mappable facies (Vavra et al., 1992). The Talang Akar Formationconsists of deltaic facies ranging from channel sandstones, delta plain shales, and channelabandonment siltstones to prodelta and delta-front shales as well as shelfal carbonates.

Each of these facies has a distinct range of displacement pressures. The following figureshows the range of seal capacities for the different facies. Shelfal carbonates and delta-front shales are excellent seals, with displacement pressures > 1000 psia (air–mercury)and capable of trapping approximately 1000–10,000 ft of oil. Delta-plain shales are rela-tively poor seals, with displacement pressures of 80–90 psia (air–mercuryº and capable oftrapping only 90–100 ft of oil.

These values place an upper limit on the amount of hydrocarbon that can be trapped.Other factors, however, limit the sealing potential of these seals. Seal risk was defined bycombining Pd with qualitative assessments of ductility, fracturing, thickness, and lateralcontinuity (Vavra et al., 1992).

Example:Ardjuna basin,Indonesia

Figure 10–49. Data from Vavra et al. (1992).

Page 73: Chap10

Evaluating Intact Top Seal • 10-73

An encouraging development is the attempt to establish a relationship between logresponse and displacement pressure (Alger et al., 1989; Shea et al., 1993; Sneider andBolger, 1993; Sneider and Neasham, 1993). Resistivity logs directly reflect the relativeproportions of clay minerals and organics to quartz and other minerals. Clay-rich sealswith high displacement pressures have low resistivities. Clay-poor rocks with lower dis-placement pressures have high resistivities (Sneider and Bolger, 1993).

Recentdevelopments

Estimating Pd from Sedimentary Facies and Well Logs, continued

A study of an African basin found differences in displacement pressure among flood plain,lacustrine, and overbank facies (Shea et al., 1993). Flood plain mudstones have displace-ment pressures of 1,500 psi (air–mercury) and can trap more than 300 m (1,000 ft) of oil.Lacustrine and overbank facies contain some siltstones that have displacement pressuresof 500 psi (air–mercury) or less. These seal–rock facies can be correlated among wells andused to predict lateral facies variations and seal distribution in uncored wells.

Example:Onshore Africanbasin

Page 74: Chap10

10-74 • Evaluating Top and Fault Seal

Displacement pressure can be estimated from pore size using Washburn’s (1921) equation,

Pd =–2γcosθ

r

Method

Estimating Pd from Pore Size

Pore throat radius, r, can be estimated in three ways:• Thin-section analysis of pore throats (Dullien and Dhawan, 1974; Etris et al., 1988;

Macdonald et al., 1986; see also Wardlaw, 1990)• Thin-section analysis of grain size and assumptions of spherical grains and rhombohe-

dral packing (Berg, 1975)• Empirical correlation of permeability, porosity, and pore throat radius (Wardlaw and

Taylor, 1976; Wells and Amaefule, 1985; Wardlaw, 1990; Pittman, 1992)

Although these methods can estimate the Pd of sands, they do not apply easily to shalesand finer grained rocks that comprise top seals. The assumptions of spherical grains andrhombohedral packing used to infer pore throat radius to not apply to shales that containplate-like clay minerals (Berg, 1975). Nor is the pore-size distribution easily determinedfrom thin sections of fine-grained shales (Krushin, 1993). Use of permeability and porosi-ty is thwarted by the lack of a distinct apex in the mercury injection data of low-perme-ability rocks (Pittman, 1992), as well as the difficulty of measuring the permeability ofshales.

Techniques forestimating r

Predicting top seal capacity is difficult—even in idealized experiments in which all vari-ables are known to a degree unobtainable in practical prospect assessment. The predictedheight of hydrocarbon columns calculated from pore throat diameters in uniform glassbead packs is 160–500T larger than heights actually observed in experiments (Catalan etal., 1992). If we cannot predict the height of hydrocarbon columns trapped in a controlledexperiment with uniform glass beads and known variables, then we must be cautious inprospect analysis.

Theory andexperiment

Page 75: Chap10

Evaluating Intact Top Seal • 10-75

A range of seal capacities have been reported for various rock types. In addition, a num-ber of other factors—depth, hydrocarbon phase, seal thickness, fault-dependent leakpoints—affect the height of trapped hydrocarbon columns.

Introduction

Subsection C3

Seal Capacity

This subsection contains the following topics.

Topic Page

Seal Capacity of Different Rock Types 10–76

Variation in Seal Capacity with Depth and Hydrocarbon Phase 10–77

Seal Capacity and Two-Phase Hydrocarbon Columns 10–79

Seal Thickness 10–80

Fault-dependent Leak Points, Continuity, and Charge 10–81

In thissubsection

Page 76: Chap10

10-76 • Evaluating Top and Fault Seal

The figure to the right shows therange of seal capacities of differ-ent rock types. This figure wascompiled from published displace-ment pressures based upon mer-cury capillary curves. Columnheights were calculated using a35°API oil at near-surface condi-tions with a density of 0.85 g/cm3,an interfacial tension of 21dynes/cm, and a brine density of1.05 g/cm3. Data were compiledfrom Smith (1966), Thomas et al.(1968), Schowalter (1979), Wellsand Amaefule (1985), Melas andFriedman (1992), Vavra et al.(1992), Boult (1993), Khrushin(1993), and Shea et al. (1993).

Range ofcapacities

Seal Capacity of Different Rock Types

These data show the following:• Good shales can trap thousands of feet of hydrocarbon column.• Most good sands can trap only 50 ft or less of oil column.• Poor sands and siltstones can trap 50–400 ft of oil column.

Generalizations

Shales have high displacement pressures and can trap large columns of oil as large as1830 m (6000 ft). Nonsmectite shales have pore throat radii of less than 12 nm and cantrap gas columns of more than 1000 m (3,000 ft) (Krushin, 1993). Shales in theCretaceous section of the Powder River basin have displacement pressures of 1000–4000psi and can trap gas columns of 460–1830 m (1500–6,000 ft) (Jiao et al., 1993). The short-est oil columns among the shale data include some true shales as well as siltstones, siltymudstones, and interbedded sand/shale cores.

Shale seals

Sands commonly have low displacement pressures and can trap only small oil columns.Three-quarters of the sands, most of which are Gulf Coast reservoirs, are capable of trap-ping less than 50 ft of oil. Sands can have sufficiently high displacement pressures to traphundreds of feet of oil. Oil column heights between 50–400 ft are from sands with diage-netic pore fillings, tight gas sands, and very fine-grained sands that probably include silt-stones.

Sand seals

Carbonates have a wide range of displacement pressures. Some carbonates can seal asmuch as 1500–6000 ft of oil. These better seals are argillaceous limestones and shelf car-bonates. In the Gulf Coast basin, shorter oil columns are sealed by grainstones, mud-stones, and wackestones of the Smackover Formation and chalk.

Carbonateseals

Figure 10–50.

Page 77: Chap10

Evaluating Intact Top Seal • 10-77

Seal capacity is controlled by the physical properties of both the seal and the hydrocarbon.Because the seal alone does not control seal capacity . . . • It is possible to trap more gas than oil despite the higher buoyant pressure of gas, • At depths below 9,000–10,000 ft a seal is always capable of trapping more gas than oil,

and • The dominance of gas in the deeper parts of basins may reflect seal capacity as well as

maturation.

Introduction

Variation in Seal Capacity with Depth and Hydrocarbon Phase

Compaction and diagenesis during burial cause a progressive reduction in pore throats inmost seal lithologies. In addition, the interfacial tension of the hydrocarbons changes withdepth and affects seal capacity. Most importantly, the interfacial tension of oil and gaschanges at different rates (Watts, 1987; Schowalter, 1979).

Changes withdepth

Because the interfacial tension ofgas increases at a different ratefrom that of oil with depth, it ispossible to trap more gas than oil.The following figure shows the sealcapacity (in feet) of a top seal fordifferent hydrocarbon compositionswith depth (after Watts, 1987).Curves show the seal capacity forthree different oils, ranging from30–40° API and a GOR of 400–800.Methane has two different pres-sure gradients (0.7 and 0.45 psi/ft.The pore throat of the seal decreas-es with depth and compaction.

Below 9000 ft, the seal capacity isgreater for gas than it is for any ofthe oil compositions. Even above9000 ft, the seal capacity for gas isgreater for some oil compositions.The seal capacity for gas is greatestat a normal pressure gradient andless for an overpressured gradient.

Seal capacityfor oil and gas

Figure 10–51. After Watts, 1987; courtesy Marine and Petroleum Geology.

Page 78: Chap10

10-78 • Evaluating Top and Fault Seal

Diffusive loss of gas through a top seal can also limit seal capacity (Leythaeuser et al.,1982; Nelson and Simmons, 1995). Seal capacity is not simply related to the displacementpressure. Although diffusion can be important in entrapment of gas, diffusion coefficientsfor oil are so small that diffusion does not affect seal capacity relative to oil.

Estimates of the diffusive loss of gas in the McClave field, Colorado, demonstrate that 57bcf of gas can diffuse from a trap in only 0.5–5.0 m.y. (Nelson and Simmons, 1995).

Gas loss bydiffusion

Variation in Seal Capacity with Depth and Hydrocarbon Phase, continued

Seal capacity rather than maturation may cause the deeper portions of some basins to begas prone. Traditionally the deep, gas-prone parts of many basins have been interpretedto be a result of increasing maturation. An alternative explanation is that seals at greaterdepth are more effective at trapping gas than oil.

Gas and oildistribution inbasins

Page 79: Chap10

Evaluating Intact Top Seal • 10-79

Seal capacity depends on both the hydrocarbon phase and the physical character of thetop seal. Since displacement pressure is a function of the seal and the coefficient of inter-facial tension of the hydrocarbon phase beneath the seal, it is possible to trap a thickertwo-phase hydrocarbon column than oil alone or gas alone (Watts, 1987). The gas in con-tact with the base of the seal determines the displacement pressure (Pd) of the seal. Thebuoyant effect of the oil column, however, is less than that of a pure gas column, and agreater total hydrocarbon column can be trapped.

Two phases are better than one

Seal Capacity and Two-Phase Hydrocarbon Columns

The following figure compares the seal capacity of a top seal with a single-phase oil accu-mulation, a single-phase gas accumulation, and a two-phase accumulation with both anoil leg and a gas cap. The largest hydrocarbon column is sealed by the two-phase accumu-lation. This best applies to traps with a geometry such that only the gas column is in con-tact with the seal. It also applies to fault traps.

Single phasevs. two phase

Figure 10–52. After Watts, 1987; courtesy Marine and Petroleum Geology.

Page 80: Chap10

10-80 • Evaluating Top and Fault Seal

There is no simple relationship between seal thickness and the height of the hydrocarboncolumn. Seals can be extremely thin—less than 1 m thick that seal individual hydrocar-bon accumulations. There are also examples of traps with thick shale seals that are dry.This difficulty in establishing a relationship between seal thickness and column is espe-cially true since many hydrocarbon columns are controlled by fault-related spill pointsthat are independent of top seal thickness.

How thick isnecessary?

Seal Thickness

Data compiled from fields in California and the Rocky Mountains show no relationshipbetween seal thickness and hydrocarbon column height (Zieglar, 1992). Nonetheless,some workers have suggested a correlation between seal thickness and seal capacity(Nederlof and Mohler, 1981; Sluijk and Nederlof, 1984).

Seal thicknessstudies

Seal thickness is not an independent variable. Thin seals have a higher probability ofbeing laterally discontinuous, of being fractured completely, or of having local variationsin fracture intensity or pore throat diameter that provide a leakage pathway. Similarly,thick seals have a higher probability of being laterally continuous, having fractures ter-minate within the seal, and having at least one shale lamina with a high displacementpressure.

Seal continuityand fracturing

Page 81: Chap10

Evaluating Intact Top Seal • 10-81

In many basins the major control on hydrocarbon column heights is not the displacementpressure of top seals but fault-dependent leak points, the lateral continuity of seals, andcharge.

Introduction

Fault-dependent Leak Points, Continuity, and Charge

A comparison of top seal capacity determined from displacement pressures and actualhydrocarbon column heights in one African basin demonstrates that while top seals arecapable of trapping more than 1000 ft of oil, most traps contain only a few hundred feet ofoil (Shea et al., 1993). Oil column heights in this basin are controlled instead by fault-dependent leak points (Allard, 1993; Shea et al., 1993).

Similarly, only 5–10% of the fields in the Rocky Mountains and California are thought tohave hydrocarbon column heights controlled by top seal capillary properties (Zieglar,1992). The remaining 90–95% are controlled by some other factor, including charge,faults, and synclinal spill points.

Fault-dependent leakpoints

The lateral extent and continuity of top seals can create and destroy plays within basins.In many basins, thick, continuous, areally extensive sequences of shale or salt act asregional top seals. The Gippsland basin and North Sea are only two of numerous exam-ples. In some basins, this regional seal is lacking or limited. In the Gulf Coast, the NorthFrisco City field exists because the Buckner Anhydrite, a regional seal, is missing above alocal basement high. Only where the regional seal is breached are hydrocarbons able toescape from the Smackover reservoir into the overlying Lower Haynesville sands(Stephenson et al., 1992). Where the Buckner Anhydrite is continuous, there is noHaynesville play.

Seal continuity

Page 82: Chap10

10-82 • Evaluating Top and Fault Seal

Although it is tantalizing to be able to measure the capillary properties of a seal and thenassess hydrocarbon volumes, some practical problems exist:• It is not easy to characterize the displacement pressure of a top seal from a few, or

many, measurements.• Few empirical field studies compare the predicted and observed hydrocarbon column

heights. This is worrisome because idealized laboratory experiments have significanterrors in the predicted hydrocarbon column heights.

• Hydrocarbon saturations required for flow through top seals—and, consequently, dis-placement pressures—may be much higher than commonly assumed.

• Hydrocarbon columns beneath breached seals and hydrocarbon-wet seals may not berelated to the capillary properties of the seal in any easily understood way.

• Diffusion can cause loss of large volumes of gas but not oil through a top seal.• Hydrodynamic flow can alter top seal capacity.

Introduction

Subsection C4

Pitfalls and Limitations of Estimating Seal Capacity

This subsection contains the following topics.

Topic Page

Difficulty of Characterizing Pd of a Seal 10–83

Does the Theory Predict Reality? 10–84

Saturations Required for Hydrocarbon Flow 10–85

Seal Capacity of Breached and Hydrocarbon-wet Seals 10–86

Hydrodynamic Flow and Pressure Transients 10–87

In thissubsection

Page 83: Chap10

Evaluating Intact Top Seal • 10-83

It is difficult to determine the displacement pressure of a top seal.• First, seal capacity is based on the presence of a single, continuous, high-Pd layer, not

upon an arithmetic average of all Pd values. The trapping capacity is determined bythe highest displacement pressure within a seal, theoretically even if that interval is alayer only one grain thick. A 500-m-thick siltstone may appear incapable of trappingmore than 1 m of hydrocarbon. The presence of a 1-cm-thick claystone layer, however,may create a seal for thousands of meters of hydrocarbon. Predicting these local sealsis difficult and adds a measure of risk to any evaluation of intact top seals using capil-lary theory.

• Second, experiments demonstrate that hydrocarbon migration through a seal does notoccur along a broad, uniform front but along narrow fingers reflecting local, tortuouspathways of low displacement pressure (Dembecki and Anderson, 1989; Catalan et al.,1992).

Reasons

Difficulty of Characterizing Pd of a Seal

These tortuous pathways develop even in bead packs of uniform, closely packed glassspheres. Natural seals are even more heterogeneous.

The difficulty of defining these narrow zones during sampling complicates our ability topredict the critical displacement pressure of a seal. Although 99% of the samples from atrap may indicate a top seal capable of trapping a 500-m oil column, the seal may wellleak through a narrow pathway only centimeters in diameter.

Pathways inexperiments

Page 84: Chap10

10-84 • Evaluating Top and Fault Seal

Rarely is the actual column of trapped hydrocarbon compared with the prediction frommeasured displacement pressures. Empirical studies are needed because simple, ideal-ized laboratory experiments show significant errors between the predicted and observedhydrocarbon column heights (Catalan et al., 1992). These experiments at best have col-umn heights 19–23% larger than predicted. At worst, they are 125–217% larger than pre-dicted. (The larger differences are probably the result of short equilibration times.)

Introduction

Does the Theory Predict Reality?

Benton field, Illinois, in one example in which the trapped column heights match themeasured displacement pressure of the top seal (Sneider and Neasham, 1993). Measureddisplacement pressures predict the top seal is capable of trapping 29–34 m (94–110 ft) ofhydrocarbon. The actual column of hydrocarbon in the field in the Tar Springs reservoir is29 m (95 ft).

Benton field

The Bodalla South field, Eromanga basin, Australia, has short oil columns—less than 12 m high—that appear to be limited by the displacement pressure of the top seal (Boult,1993). The following cross section shows the top seal and oil accumulations. The top seal,the Birkhead Formation, consists of fluvial channel and point bar deposits as well asmore shale-prone levee bank, crevasse splay, floodplain, and coal swamp facies. Measureddisplacement pressures of the top seal have an average range of 150–200 psi (air-mer-cury) and can seal 10.86–12.64 m of oil. The maximum height of the actual trapped oil col-umn is 11.25 m.

Bodalla Southfield

Although predicted and observed data apparently agree, units within the top seal may beable to trap much greater oil columns. Measured displacement pressures of the shalyoverbank facies are greater than 3000 psi (air–mercury). These facies are not thought tobe effective seals because they are either thin or lack lateral continuity (Boult, 1993).More empirical studies are needed.

Top seal qualityvariation

Figure 10–53. After Boult, 1993; courtesy Marine and Petroleum Geology.

Page 85: Chap10

Evaluating Intact Top Seal • 10-85

Hydrocarbons flow through a water-wet seal when there is a continuous, interconnectedpathway of hydrocarbon-filled pore space. Flow through a seal occurs with saturations of4.5–17% of the rock pore volume, averaging 10% (Schowalter, 1979). This is one reasondisplacement pressure is defined at 10% saturation.

Traditionaldefinitions

Saturations Required for Hydrocarbon Flow

Other experiments, including lower porosity rocks, indicate much higher oil saturations of25–91% may be required (England et al., 1987). The current assumption that seal capaci-ty is based upon the displacement pressure at only 10% saturation could be very mislead-ing. The value could be two or more orders of magnitude greater than that predicted at10% saturation.

Possiblemodifications

Page 86: Chap10

10-86 • Evaluating Top and Fault Seal

Hydrocarbon column heights may not be related only to the displacement pressure of thetop seal. Once a seal has been breached and hydrocarbons forced through the seal, theo-retically the hydrocarbon column will shrink until the buoyant pressure equals the dis-placement pressure and the system again seals. In practice, however, hydrocarbons con-tinue to flow through the seal until there is no longer a continuous hydrocarbon filament.Although the process is not completely understood, laboratory studies suggest that flowcontinues until the hydrocarbon column shrinks to half its original height (Roof, 1970;Schowalter, 1979).

Introduction

Seal Capacity of Breached and Hydrocarbon-wet Seals

The preceding discussion suggests that in basins where there is charge sufficient to fill alltraps to maximum seal capacity, traps limited by the capillary properties of intact topseals should be half full (Watts, 1987). Rather than the hydrocarbon columns matchingthe displacement pressures of the seal, all filled traps would have leaked hydrocarbonsuntil the buoyant pressure (Pb) is half the displacement pressure (Pd). Continued chargingof traps after initial fill could result in larger hydrocarbon columns. Only in basins withinsufficient hydrocarbons to fill all traps to the maximum seal capacity, or with chargeafter leakage, would there be a large number of traps greater than half full.

The following figure illustrates the fill history of a trap with some finite seal capacity thathas been filled to seal capacity and then leaked. The final hydrocarbon column is half ofthat predicted by the displacement pressure of the top seal.

Most traps halffull

Figure 10–54. After Boult, 1993; courtesy Marine and Petroleum Geology.

Estimates of seal capacity from measured displacement pressures commonly assume theseal is water wet. Oil-wet seals may be more common than we think. Organic-rich shales,a common top seal, are probably oil wet (Cuiec, 1987). Similarly, episodic leakage ofhydrocarbons through a seal may alter the seal capacity.

Water wet vs. oil wet

Page 87: Chap10

Evaluating Intact Top Seal • 10-87

Pressure gradients and the resulting buoyant pressures are not always static. Bothhydrodynamic flow and pressure transients change seal capacity.

Introduction

Hydrodynamic Flow and Pressure Transients

Fluid pressure gradients may fluctuate dramatically during faulting and basin evolution(Sibson et al., 1975). Estimates of trapping capacity based upon the capillary modelassume a static pressure gradient or a uniform regional hydrodynamic gradient.Measurements of fluid inclusions, however, suggest pressure transients along faults of asmuch as 126 MPa (1,825 psi) (Parry and Bruhn, 1990). Similar episodic fluid flow eventsare inferred from sandstone cements in the North Sea (Robinson and Gluyas, 1992).Leakage through seals and seal capacity may be as episodic as hydrocarbon generation,migration, and pressure transients.

Pressuretransients

A hydrodynamic gradient will either increase or decrease the height of a trapped hydro-carbon column (Schowalter, 1979; Dahlberg, 1982; Lerche and Thomsen, 1994). Flow inthe direction of the buoyant vector decreases the seal capacity. Flow opposite the directionof the buoyant vector increases the seal capacity.

The hydrodynamic effect has been demonstrated to be important in trapping hydrocar-bons in the western Canada basin (Dahlberg, 1982; Lerche and Thomsen, 1994).

Hydrodynamicflow

Page 88: Chap10

10-88 • Evaluating Top and Fault Seal

Alger, R.P., D.L. Luffel, and R.B. Truman, 1989, New unified method of integrating corecapillary pressure data with well logs: Society of Petroleum Formation Evaluation, vol. 4,no. 2, p. 145–152.

Allan, U.S., 1989, Model for hydrocarbon migration and entrapment within faulted struc-tures: AAPG Bulletin, vol. 72, no. 7, p. 803-811.

Allard, D.M., 1993, Fault leak controlled trap fill, rift basin examples (abs.), in J. Ebanks,J. Kaldi, and C. Vavra, eds., Seals and Traps: A Multidisciplinary Approach: AAPGHedberg Conference, Crested Butte, Colorado, June 21–23.

Bell, J.S., 1990, Investigating stress regimes in sedimentary basins using informationfrom oil industry wireline logs and drilling records, in A. Hurst, M.A. Lovell, and A.C.Morton, eds., Geological Applications of Wireline Logs: Geological Society London SpecialPublication 48, p. 305–325.

Berg, R.R., 1975, Capillary pressure in stratigraphic traps: AAPG Bulletin, vol. 59, no. 6,p. 939–956.

Boult, P.J., 1993, Membrane seal and tertiary migration pathways in the Bodalla Southoilfield, Eronmanga Basin, Australia: Marine and Petroleum Geology, vol. 10, no. 1, p.3–13.

Bouvier, J.D., C.H. Kaars-Sijpesteigen, D.F. Kluesner, C.C. Onyejekwe, and R.C. VanderPal, 1989, Three-dimensional seismic interpretation and fault sealing investigations, NunRiver field, Nigeria: AAPG Bulletin, vol. 73, no. 11, p. 1397–1414.

Breckles, I.M., and H.A.M. Van Eekelen, 1982, Relationship between horizontal stressand depth in sedimentary basins: Journal of Petroleum Technology, vol. 34, no. 9, p.2191–2199.

Brennan, R.M., and M.R. Annis, 1984, A new fracture gradient prediction technique thatshows good results in Gulf of Mexico abnormal pressure: SPE paper 13210, 6 p.

Buck, S., and G. Robertson, 1996, Fault seal behavior at Beryl field, UK North Sea: obser-vations from 20 years of production, drilling and injection data: AAPG Annual MeetingAbstracts, San Diego, May 19–22, p. A20.

Caillet, G., 1993, The caprock of the Snorre field (Norway): a possible leakage byhydraulic fracturing: Marine and Petroleum Geology, vol. 10, no. 1, p. 42–50.

Catalan, L. F. Xiaown, I. Chatzis, and F.A.L. Dullien, 1992, An experimental study of sec-ondary oil migration: AAPG Bulletin, vol. 76, no. 5, p. 638–650.

Chong, K.P., P.M. Hoyt, J.W. Smith, and B.Y. Paulsen, 1980, Effects of strain rate on oilshale fracturing: International Journal of Rock Mechanics, vol. 17, no. 1, p. 35–43.

Corbett, K., M. Friedman, and J. Spang, 1987, Fracture development and mechanicalstratigraphy of Austin Chalk, Texas: AAPG Bulletin, vol. 71, no. 1, p. 17–28.

Section D

References

Page 89: Chap10

References • 10-89

Cuiec, L., 1987, Wettability and oil reservoirs, in J. Kleppe, E.W. Berg, A.T. Buller, O. Hjemeland, and O. Torsaeter, eds., North Sea Oil and Gas Reservoirs: London,Graham and Trotman, p. 193–207.

Dahlberg, E.C., 1982, Applied Hydrodynamics in Petroleum Exploration: New York,Springer-Verlag, 161 p.

Dembecki, H., Jr., and M.J. Anderson, 1989, Secondary migration of oil: experiments sup-porting efficient movement of separate, buoyant oil phase along limited conduits: AAPGBulletin, vol. 73, no. 8, p. 1018–1021.

Doveton, J.H., 1986, Log analysis of subsurface geology: New York, John Wiley & Sons,273 p.

Downey, M.W., 1984, Evaluating seals for hydrocarbon accumulations: AAPG Bulletin,vol. 68, no. 11, p. 1752–1763.

Dullien, F.A.L., and G.K. Dhawan, 1974, Characterization of pore structure by a combina-tion of quantitative photomicrography and mercury porosimetry: Journal of Colloid andInterface Science, vol. 47, no. 2, p. 337–349.

Dutta, N.C., 1987, Fluid flow in low permeability porous media, in B. Doligez, ed.,Migration of Hydrocarbons in Sedimentary Basins: Paris, Editions Technip, p. 567–596.

Earlougher, R.C., Jr., and H. Kazemi, 1980, Practicalities of detecting faults from builduptesting: JPT, January 1980, p. 18–20.

Eaton, B.A., 1969, Fracture gradient prediction and its application in oilfield operations:Trans. AIME, October, p. 1353–1360.

Eisenberg, R.A., R.J. Brenneman, and A.A. Adeogba, 1996, Integrated fault seal analysisand risk assessment: Okan and Meren fields, Nigeria: AAPG Annual Meeting Abstracts,vol. 5, San Diego, May 19–22, p. 41.

Engelder, T., and A. Lacazette, 1990, Natural hydraulic fracturing, in N. Barton and O.Stephansson, eds., Rock Joints: Rotterdam, A.A. Balkema, p. 35–43.

——— and J. Leftwich, 1993, An analysis of geopressure profiles from south Texas: thesearch for higher quality reservoirs in the Tertiary sections of the Gulf Coast: AAPGHedberg Research Conference, unpublished abstract.

England, W.A., A.S. Mackenzie, D.M. Mann, and T.M. Quickley, 1987, The movement andentrapment of petroleum fluids in the subsurface: Journal of the Geological Society,London, vol. 144, no. 2, p. 327–347.

Evans, C.J., and N.R. Brereton, 1990, In situ crustal stress in the United Kingdom fromborehole breakouts, in A. Hurst, ed., Geological Applications of Wireline Logs: Geologicalsociety Special Publication 48, p. 327–338.

Evans, D.M., 1966, The Denver area earthquakes and the Rocky Mountain Arsenal dis-posal well: The Mountain Geologist, vol. 3, no. 1, p. 23–36.

References, continued

Page 90: Chap10

10-90 • Evaluating Top and Fault Seal

Fertl, W.H., and W.G. Leach, 1988, Economics of hydrocarbon reserves in overpressuredreservoirs below 18,000 feet in south Louisiana: SPE paper 18146, 16 p.

Firoozabadi, A., and H.J. Ramey, Jr., 1988, Surface tension of water-hydrocarbon systemsat reservoir conditions: Journal of Canadian Petroleum Technology, vol. 27, no. 3, p.41–48.

Gibson, R.G., 1994, Fault-zone seals in siliciclastic strata of the Columbus basin, offshoreTrinidad: AAPG Bulletin, vol. 78, no. 9, p. 1372–1385.

Grunau, H.R., 1987, A world-wide look at the cap-rock problem: Journal of PetroleumGeology, vol. 10, no. 3, p. 245–266.

Harding, T.P., and A.C. Tuminas, 1988, Interpretation of footwall (lowside) fault trapssealed by reverse faults and convergent wrench faults: AAPG Bulletin, vol. 72, no. 7, p.738–757.

Hardman, R.F.P., and J.E. Booth, 1989, Structural interpretation of hydrocarbon trapssealed by basement normal fault block faults at stable flank of foredeep basins and at riftbasins: AAPG Bulletin, vol. 73, no. 7, p. 813–840.

——— and ———, 1991, The significance of normal faults in the exploration and produc-tion of North Sea hydrocarbons, in A.M. Roberts, G. Yielding, and B. Freeman, eds., TheGeometry of Normal Faults: London, Geological Society of London, p. 1–16.

Hoshino, K., H. Koide, K. Inami, S. Iwamura, and S. Mitsui, 1972, Mechanical propertiesof Japanese Tertiary sedimentary rocks under high confining pressures: GeologicalSurvey of Japan Report 244, 200 p.

Hubbert, M.K., 1953, Entrapment of petroleum under hydrodynamic conditions: AAPGBulletin, vol. 37, no. 8, p. 1954–2026.

——— and D.G. Willis, 1957, Mechanics of hydraulic fracturing: JPT, vol. 9, no. 6, p.153–168.

Jev, B.I., C.H. Kaars-Sijpesteign, M.P.A.M. Peters, N.W. Watts, and J.T. Wilkie, 1993,Akaso field, Nigeria: use of integrated 3-D seismic, fault-slicing, clay smearing and RFTpressure data on fault trapping and dynamic leakage: AAPG Bulletin, vol. 77, no. 8, p.1389–1404.

Jiao, Z.D., R.C. Surdam, R.S. Martinsen, and W.P. Iverson, 1993, Capillary seals and pres-sure compartment boundaries in the Cretaceous section in the Power River basin, in J.Ebanks, J. Kaldi, and C. Vavra, eds., Seals and Traps: A Multidisciplinary Approach:AAPG Hedberg Research Conference, unpublished abstract.

Katz, A.J., and A.H. Thompson, 1986, Quantitative prediction of permeability in porousrock: Physical Review Bulletin, vol. 34, p. 8179–8181.

Knutson, C.A., and R. Erga, 1991, Effect of horizontal and vertical permeability restric-tions in the Beryl reservoir: Journal of Petroleum Technology, vol. 43, p. 1502–1509.

References, continued

Page 91: Chap10

References • 10-91

Krushin, J., 1993, Entry pore throat size of nonsmectite shales, in J. Ebanks, J. Kaldi,and C. Vavra, eds., Seals and Traps: A Multidisciplinary Approach: AAPG HedbergResearch Conference, unpublished abstract.

Lama, R.D., and V.S. Vutukuri, 1978, Handbook of Mechanical Properties of Rocks:Rockport, MA, Trans. Technical Publications.

Leach, W.G., 1993a, Fluid migration, HC concentration in south Louisiana Tertiary sands:Oil & Gas Journal, vol. 91, no. 11, p. 71–74.

———, 1993b, Maximum hydrocarbon window determination in south Louisiana: Oil &Gas Journal, vol. 91, no. 13, p. 81–84.

Leith, T.L., I. Kaarshad, J. Connan, J. Pierron, and G. Caillet, 1993, Recognition ofcaprock leakage in the Snorre field, Norwegian North Sea: Marine and PetroleumGeology, vol. 10, no. 1, p. 29–41.

Lerche, I., 1990, Basin Analysis I: San Diego, Academic Press, 562 p.

——— and R.O. Thomsen, 1994, Hydrodynamics of Oil and Gas: New York, PlenumPress, 308 p.

Leythauser, D., R.G. Schaeffer, and A. Yukler, 1982, Role of diffusion in primary migrationof hydrocarbons: AAPG Bulletin, vol. 66, no. 4, p. 408-–429.

Lorenz, J.C., L.W. Teufel, and N.R. Warpinski, 1991, Regional fractures: a mechanism forthe formation of regional fractures at depth in flat-lying reservoirs: AAPG Bulletin, vol.75, no. 11, p. 1714–1737.

Macdonald, I.F., P. Kaufmann, and F.A.L. Dullien, 1986, Quantitative image analysis offinite porous media 2: Specific genus of cubic lattice models and Berea sandstone: J.Microscopy, vol. 144, p. 297–316.

Mann, D.M., and A.S. Mackenzie, 1990, Prediction of pore fluid pressures in sedimentarybasins: Marine and Petroleum Geology, vol. 7, no. 1, p. 55–65.

Matthews, W.R., and J. Kelly, 1967, How to predict formation pressure and fracture gradi-ent from electric and sonic logs: Oil & Gas Journal, vol. 65, no. 8, p. 92–106.

Melas, F.F., and G.M. Friedman, 1992, Petrophysical characteristics of the JurassicSmackover Formation, Jay field, Conecuh Embayment, Alabama and Florida: AAPGBulletin, vol. 76, no. 1, p. 81–100.

Myers, J.D., 1968, Differential pressures: a trapping mechanism in Gulf Coast oil and gasfields: Gulf Coast Assoc. of Geologists Transactions, vol. 18, p. 56–80.

Nederlof, M.N., and H.P. Mohler, 1981, Quantitative investigation of trapping effect ofunfaulted caprock: AAPG Bulletin, vol. 65, no. 6, p. 964.

Nelson, J.S., and E.C. Simmons, 1995, Diffusion of methane and ethane through thereservoir cap rock: Implications for the timing and duration of catagenesis: AAPGBulletin, vol. 79, no. 7, p. 1064–1074.

References, continued

Page 92: Chap10

10-92 • Evaluating Top and Fault Seal

Niemann, J., and M. Krowlow, 1992, Delineation of a pressure fault seal from shale resis-tivities, in J. Ebanks, J. Kaldi, and C. Vavra, eds., Seals and Traps: A MultidisciplinaryApproach: AAPG Hedberg Research Conference, unpublished abstract.

Parry, W.T., and R.L. Bruhn, 1990, Fluid pressure transients on seismogenic normalfaults: Tectonophysics, vol. 179, no. 3–4, p. 335–344.

Prasad, R.K., 1975, Pressure transient analysis in the presence of two intersecting bound-aries: Society of Petroleum Engineers paper 4560.

Pittman, E.D., 1992, Relationship of porosity and permeability to various parametersderived from mercury injection-capillary pressure curves for sandstone: AAPG Bulletin,vol. 76, no. 2, p. 191–198.

Robinson, A., and J. Gluyas, 1992, Duration of quartz cementation in sandstones, NorthSea and Haltenbanken basins: Marine and Petroleum Geology, vol. 9, no. 3, p. 324–327.

Roof, J.G., 1970, Snap-off of oil droplets in water-wet pores: SPE Journal, vol. 10, no. 1, p.85–90.

Scharr, G., 1976, The occurrence of hydrocarbons in overpressured reservoirs of theBaram delta, offshore Sarawak, Malaysia: Fifth Annual Convention, Indonesian Petro-leum Association, Proceedings, p. 163–169.

Schowalter, T.T., 1979, Mechanics of secondary hydrocarbon migration and entrapment:AAPG Bulletin, vol. 63, no. 5, p. 723–760.

Shea, W.T., J.R., Schwalbach, and D.M. Allard, 1993, Integrated rock-log evaluation offluvio-lacustrine seals, in J. Ebanks, J. Kaldi, and C. Vavra, eds., Seals and Traps: AMultidisciplinary Approach: AAPG Hedberg Research Conference, unpublished abstract.

Sibson, R.H., J. Moore, and A.H. Rankin, 1975, Seismic pumping—a hydrothermal fluidtransport mechanism: Journal of the Geological Society of London, vol. 131, p. 653–659.

Skerlec, G.M., 1982, Risking top seals in the Central Graben: Exxon Production ResearchCompany internal report.

———, 1990, SEALS: A short course for risking top seal and fault seal: Franklin,Pennsylvania, SEALS International, 600 p.

———, 1992, Snap, crackle & pot: risking top seal integrity: AAPG Annual Conventionprogram abstracts, p. 21.

———, 1997a, Atlas of fault seal behavior in the North Sea/Norwegian Sea: Franklin,Pennsylvania, SEALS International, 3948 p.

———, 1997b, Atlas of fault seal behavior in the Gulf Coast: Franklin, Pennsylvania,SEALS International, 4356 p.

Slujik, D., and M.H. Nederlof, 1984, Worldwide geological experience as a systematicbasis for prospect appraisal, in G. Demaison and R.J. Murris, eds., PetroleumGeochemistry and Basin Evaluation: AAPG Memoir 35, p. 15–26.

References, continued

Page 93: Chap10

References • 10-93

Smith, D.A., 1966, Theoretical considerations of sealing and non-sealing faults: AAPGBulletin, vol. 50, no. 2, p. 363–374.

———, 1980, Sealing and non-sealing faults in the Gulf Coast Salt basin: AAPG Bulletin,vol. 64, no. 2, p. 145–172.

Sneider, R.M., and G. Bolger, 1993, Estimating seals from wireline logs of clastic sealsand reservoir intervals, in J. Ebanks, J. Kaldi, and C. Vavra, eds., Seals and Traps: AMultidisciplinary Approach: AAPG Hedberg Research Conference, unpublished abstract.

———, and J. Neasham, 1993, Comparison of seal capacity determination (SCD) fromcores with SCD’s from cuttings of the same interval, in J. Ebanks, J. Kaldi, and C. Vavra,eds., Seals and Traps: A Multidisciplinary Approach: AAPG Hedberg ResearchConference, unpublished abstract.

Stearns, D.W., and M. Friedman, 1972, Reservoirs in fractured rock: AAPG Memoir 16, p.82–106.

Stephenson, M., J. Cox, and P. Jones-Fuentes, 1992, How 3D seismic-CAEX combinationaffected development of N. Frisco City field in Alabama: Oil & Gas Journal, vol. 90, no. 43,p. 127–130.

Stewart, G., A. Gupta, and P. Westaway, 1984, The interpretation of interference tests in areservoir with sealing and partially communicating faults: Society of Petroleum Engi-neers, paper 12967.

Thomas, L.K., D.L. Katz, and M.R. Ted, 1968, Threshold pressure phenomena in porousmedia: Transactions of SPE, vol. 243, p. 174–184.

Ungerer, P. J. Burrus, B. Doligez, P.Y. Chenet, and F. Bessis, 1990, Basin evaluation byintegrated two-dimensional modeling of heat transfer, fluid flow, hydrocarbon generation,and migration: AAPG Bulletin, vol., 74, no. 3, p. 309–335.

van Poollen, H.K., 1965, Drawdown curves give angle between intersecting faults: Oil &Gas Journal, vol. 63, no. 52, p. 71–75.

Van Wijhe, D.H., M. Lutz, and J.P.H. Kaasschieter, 1980, The Rotliegend in theNetherlands and its gas accumulations: Geologie en Minjbouw, vol. 59, no. 1, p. 3–24.

Vavra, C.L., J.G., Kaldi, and R.M. Sneider, 1992, Geological applications of capillary pres-sure: a review: AAPG Bulletin, vol. 76, no. 6, p. 840–850.

Wardlaw, N.C., 1990, Quantitative determination of pore structure and application tofluid displacement in reservoir rocks, in J. Kleppe, E.W. Berg, A.T. Buller, O. Hjemeland,and O. Torsaeter, eds., North Sea Oil and Gas Reservoirs: London, Graham & Trotman, p.229–243.

——— and R.P. Taylor, 1976, Mercury capillary pressure curves and the interpretation ofpore structure and capillary behavior in reservoir rocks: Canadian Petroleum GeologyBulletin, vol. 24, no. 2, p. 225–262.

References, continued

Page 94: Chap10

10-94 • Evaluating Top and Fault Seal

Washburn, E.W., 1921, Note on a method of determining the distribution of pore sizes in aporous material: Proceedings of the National Academy of Science, vol. 7, p. 115–116.

Watts, N.L., 1983, Microfractures in chalks of the Albuskjell field, Norwegian sector,North Sea: possible origin and distribution: AAPG Bulletin, vol. 67, no. 2, p. 201–234.

———, 1987, Theoretical aspects of cap-rock and fault seals for single and two-phasehydrocarbon columns: Marine and Petroleum Geology, vol. 4, no. 4, p. 274–307.

Weber, K.J., G. Mandl, W.F. Pilaar, F. Lehner, and R.G. Precious, 1978, The role of faultsin hydrocarbon migration and trapping Nigerian growth fault structures: Proceedings,Offshore Technology Conference, Houston, p. 2643–2652.

Wells, J.D., and J.O. Amafuele, 1985, Capillary pressure and permeability relationshipsin tight gas sands: SPE/DOE paper 13879.

Wilkinson, E.R., 1971, California offshore oil and gas seeps: California Oil Fields–Summary Operations, vol. 57, no. 1, p. 5–28.

Yaxley, L.M., 1987, Effect of a partially communicating fault on transient pressure behav-ior: Society of Petroleum Engineers, paper 14311.

Yielding, G., B. Freeman, and D.T. Needham, 1997, Quantitative fault seal prediction:AAPG Bulletin, vol. 81, no. 6, p. 897–917.

Zieglar, D.M., 1992, Hydrocarbon columns, buoyancy pressures, and seal efficiency: com-parisons of oil and gas accumulations in California and the Rocky Mountain area: AAPGBulletin, vol. 76, no. 4, p. 501–508.

References, continued

Many of these ideas and techniques have evolved over the past 20 years in courses andworkshops on seal analysis as well as prospect and play assessments for numerous com-panies. I am grateful to them for contributing to our understanding of seals. I especiallywish to thank my former colleagues at Exxon Production Research Company who havebeen leaders in developing many of these techniques for fault seal analysis. R. Vierbuchenfirst pointed out the relationship between shale density and ductility. Esso Explorationand Production kindly authorized permission to publish the examples of seal analysis inthe North Sea. Chevron USA provided data for the Gulf Coast field example of routinefault seal analysis.

Acknowledgments