Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

SPE 166142

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


Environmental Risk Arising From Well Construction Failure: Difference
Between Barrier and Well Failure, and Estimates of Failure Frequency
Across Common Well Types, Locations and Well Age
George E. King, Apache Corporation
Daniel E. King, WG Consulting

Copyright 2013, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Annual Technical Conference and Exhibition held in New Orleans, Louisiana, USA, 30 September–2 October 2013.

This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents of the paper have not been
reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect any position of the Society of Petroleum Engineers, its
officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written consent of the Society of Petroleum Engineers is prohibited. Permission to
reproduce in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract
Do oil and gas wells leak to the environment? The great majority of wells do not pollute. The purpose of this paper is to
explain basic concepts of well construction and illustrate differences between single barrier failure in multiple barrier well
design and outright well integrity failure that could lead to pollution, using published investigations and reviews from data sets
of over 600,000 wells worldwide. For US wells, while individual barrier failures (containment maintained and no pollution
indicated) in a specific well group may range from very low to several percent (depending on geographical area, operator, era,
well type and maintenance quality), actual well integrity failures are very rare. Well integrity failure is where all barriers fail
and a leak is possible. True well integrity failure rates are two to three orders of magnitude lower than single barrier failure
rates.

When a series of barriers fail and a leak path is formed, gas is the most common fluid lost. Common leak points are failed
gaskets or valves at the surface and are easily and quickly repaired. If the failure is subsurface, an outward leak is uncommon
due to lower pressure gradient in the well than in outside formations. Subsurface leaks in oil wells are rare and are routinely
exterior formation salt water leaking into the well towards the lower pressure in the well.

Failure frequency numbers are estimated for wells in several specific sets of environmental conditions (location, geologic
strata, produced fluid composition, soils, etc.). Accuracy of these numbers depends on a sufficient database of wells with
documented failures, divided into: 1) barrier failures in a multiple barrier system that do not create pollution, and 2) well
integrity failures that create a leak path, whether or not pollution is created. Estimated failure frequency is only for a specific
set of wells operating under the same conditions with similar design and construction quality. Well age and era of construction
are variables. There is absolutely no one-size-fits-all well failure frequency.

Introduction
This paper is a continuation of the work on environmental risk of well development that stated with a study of environmental
risk from fracturing-related activity, where potential of possible pollution from materials transport was identified, as was a
much lesser impact from leaks in the finished well (King, 2012). Environmental risk to ground water from the integrity of
producing wells is addressed in this study, which also examines several other possibilities of environmental impact during the
producing lifespan, from the end of construction up to plug and abandonment. For purposes of focus and brevity, this work is
limited to the failure potential of the constructed barriers remaining in the producing well after drilling; e.g., casing, cement,
packers, tubing, wellheads and other downhole equipment that remains part of the producing well at the handover from drilling
to production operations. The reader is referred to studies of well drilling for the mechanical equipment barriers used in
drilling (Handal, 2013; Steinsvik, 2008; Aadnoy, 2008), and to studies on injection wells for information on UIC-II injection
wells (US 40 CFR 146), and Plugging and Abandonment studies (NPC, 2011; Khalifeh, 2013; Faul, 1999). These subjects will
not be a part of this work.

The information is from government, academic and industry reports and care is taken to document the sources of all data. No
2 SPE 166142

company comparisons are made. Most information in the report is from North America as a result of the large well population
and the wealth of available data from government, academic and industry sources. Other production well data sources have
been considered to expand the analyses (Bonn, 1998; Bazzari, 1989; Liang, 2012; Rao, 2012; Sintef, 2010).

Well Design Overview – Establishing Redundant Barriers


Barriers may be active, passive, or in some cases, reactive. Active barriers such as valves can enable or prevent flow, while
passive barriers are fixed structures such as casing and cement. When barriers are used in series (nested one-inside-the-other),
a multiple barrier system is created; essentially a “defense-in-depth” barrier system (Hollnagel, 2004; Fleming, 2002; Sklet,
2006). Reactive barriers are invisible or unobtrusive in normal operations, but deploy a containment response when a pressure,

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


flow rate or other behavior limit is exceeded. In the oil and gas industry a reactive barrier may be a human or mechanical
response to an activating or triggering event.

There are four steps in developing a mechanical reactive barrier for a well (adapted from Hale, 2004):
 Definition or specification of a barrier, understanding of causes of failure, and what signals could be monitored to
help predict or sense a failure.
 A detection mechanism suitable to perceive a present or, preferably, an incipient failure,
 An activation mechanism that will alarm, notify or automatically begin a response,
 A response mechanism that is suitable to prevent, control or mitigate the failure.

The difference between drilling and production well barriers is that most production well barriers are static (available
continuously over an extended period of time, usually without requiring human observation or action) whereas most drilling
and completion activity barriers are dynamic (control is variable with time and activity). Production barriers require less
continuous monitoring compared drilling and completion barriers that are dependent on correct human activity.

In most well configurations, uncemented sections of inner pipe strings are designed to collapse under any over-pressuring
external load in the annulus before the pipe that forms the outer wall of the annulus can burst. This type of reactive barrier
protects the integrity of the outer string with a sacrificial collapse of the inner string. While effective, this fixed design is
susceptible to changes in how the well is operated, such as converting a primary flowing well to a gas lift with high pressure
lift gas in the innermost annulus. Burst strength of pipe is further reinforced by cement on the outside, although this added
burst strength is rarely considered in the typical conservative design (Zinkham, 1962). Redundant, sequential design elements
control failure-producing forces so a well can survive without loss of well integrity, even with the loss of one or more barriers
in a multi-barrier system. This contained failure approach, which is basically an over-design, is the core of engineering design
where safety is paramount. Contained failure approaches and redundant barrier systems are common in all risk-based
industries. For a well to pollute, a leak path must form and extend from the inner hydrocarbon flow path to the outside
environment. Any barrier that interrupts this flow path and prevents formation of a leak path is effective in preventing
pollution.

Barriers and Well Integrity


Oil and gas producing wells are a nested collection of pipe, cement, seals and valves that form multiple barriers between
produced well fluids and the outside environment.
Definition of barrier and integrity in well design technology as loosely defined as follows:

 Barriers are containment elements that can withstand a specific design load. These may be pipe that is effectively
cemented, seals, pipe body, valves, pressure rated housings, etc., Table 1.
 Multiple barriers are nested individual barriers designed and built to withstand a specific load without help from other
barriers. If an inside (or outside) barrier fails, the next barrier will provide isolation so that a leak path will not form.
In modern designs, the number of barriers used is typically in proportion to the hazard potential in specific areas of
the well, Figure 1. When a barrier does fail, an assessment of the leak and the integrity risk will establish whether the
leak should be repaired immediately or if the well can safely be produced until maintenance or repair can be
scheduled. State and Federal regulations cover exemptions that may or may not be granted to continue to operate
(Bourgoyne, 1999, 2000; Corneliussen, 2007; Anders, 2008; Browning, 1993; Vignes, 2008 & 2011; Anders, 2008;
Browning, 1993; Calosa, 2010; Crow, 2006; D’Alesio, 2011; Dethlefs, 2012; Dugay, 2012; Nygaard, 2010; Randohl,
2008).
 Monitored barriers may have fluid-filled gaps (annuli) between barriers that can be monitored for pressure increase or
fluid invasion. Annular gaps that are cemented to the surface are usually not monitored.
 Well integrity failure is an undesired result where all barriers in a sequence fail in such a way that a leak path is
created. Whether or not pollution occurs; however; depends on the direction of pressure differential. Since
subsurface pressure in many wells is actually lower inside the well than outside, the leak path is often into the well
and environmental pollution potential is minor or absent.
SPE 166142 3

Table 1. Examples of operations barriers:


Barrier Number Press Rating Durability of Barrier
Description Present of Producing Elements
Well Barriers
Casing + 2 to 7 Very High Very High – many
Cement cases of 70+ yr

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


producing life.
Hangers + 2 to 4 High High –internal seals
Seals
Pipe Between 1 to 3 High Very High where
Seals corrosion maint.
Packoffs and 1 to 2 Moderate Moderate – internal
Some Plugs seal.
Safety Valves 1 Moderate to Moderate – internal
(surface, tbg High seal – frequently
or annular) tested.
Valves, 4 to Moderate to Moderate – can be
Spools, seals 20+, High isolated quickly
in Prod. Tree

Figure 1 Examples of Operational Barriers


4 SPE 166142

Risk
The definition of risk used here includes the recognition that, although there is a degree of risk in every action, the frequency
of occurrence and the impact of a detrimental outcome create a risk or threat level that we can understand and accept or reject
based on what we believe, hopefully from assessment of facts. When solid occurrence numbers are not available, probability
is used as a proxy. While the use of a proxy like probability is necessary in many cases, an element of uncertainly is
inescapably included.

Wells are designed and built as pressure vessels using exact data on as many variables of the formations and producing

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


conditions as we know and considering how they will change as subterranean forces are altered by producing or injecting
fluids into rocks that have reached equilibrium in nature. One challenge to well design is that every inch of a depositional
formation is different from the inch above and the inch below; hence the need to design for the unknown and the worst load.
Well design is a geo-mechanical, fit-for-purpose engineering effort and definitely not a one-size-fits-all approach.

All phases of the well design must consider loads and forces placed
on the well from the first cementing operation through fracturing and
to the end of production. Well failure causes include simple, one-
variable induced failures and more complex failure scenarios.
Whenever possible, a failure should be traced back to a root cause.
Although examining surface failures is a practical approach (in which
pollution can be more quickly and unambiguously documented),
determining root cause of subsurface failures, where the failed
equipment cannot always be retrieved, is more difficult and the direct
or indirect environmental damage, if it occurs at all, may not be seen
for months after the incident. The most important element of risk
control is to prevent barrier failure by predicting performance of
barriers in any operating condition.

Neither human nor natural endeavors are risk free (Ritchie, 2013). To
work with risk requires an assessment of the impact and the
probability of an undesirable outcome. For this reason, the actual
expression of risk must be made based on a quantitative risk
assessment (QRA) and may be compared to other industries where
significant risk is an issue (Oakley, 2005; Martland, 2012). The
concept of ALARP (As Low As Reasonably Practicable) is widely
accepted in risk-based industries and by the public, although many
have not heard the term, Figure 2.

For example, actuarial tables of the life insurance industry on pilots


and on-board staff (showing no elevated life risk from flying in
scheduled airlines) and the public’s acceptance of the airlines as a
safe way to travel are an educated acceptance of risk. The ALARP
term comes from UK and North Sea based safety practices and law in
the area of safety-critical systems. The basic principle is that the
residual risk shall be as low as reasonably practicable, but all
endeavors accept a risk threshold that can be described as judgment of
the balance of risk and social benefit. Within that definition; however; the implied responsibility is that future risk must always
be further diminished by applying learnings and developments from study of past operations.

Possibility versus Probability


Failure frequency, impact assessment and risk rankings are needed to learn what problems are most important and require
rapid attention. Identifying problems without ranking them for frequency and impact is somewhat similar to comparing a large
asteroid collision with a tripping hazard created by a wrinkle in the carpet. Both are hazards, but one is catastrophic with a
miniscule chance of occurring and few ways to avoid it; while the other, although frequently encountered, is a minor issue and
easily corrected. Impact and frequency are used in every risk-based industry as the basis for preparation, plans, and changes to
designs.

Effectiveness of Barriers
The only constant in the geology of the earth is change, as attested to by millions of years of earthquakes, volcanic activity,
formation flows, mountain erosion, natural seeps of oil, gas and salt water and a thousand other continuously changing
SPE 166142 5

processes large and small. The natural containment barriers in reservoirs are extremely strong; proven by the fact that low
density oil and gas remained trapped there even after millions of years of major earthquakes and other tectonic evens. Some
areas, however, such as the northeastern US, have natural formation arrangements that allow slow movement of oil, gas and
salt water to the surface. These natural pathways (seeps) have been functioning for millions of years and communities of flora
and fauna have developed to make use of the output of these natural hydrocarbon seeps.

From the study of forces of ruin (wear, corrosion, erosion, decomposition, weather, cyclic loads, etc.) that degrade all things
natural and manmade; engineers describe behaviors that will destruct, and counter-actions that will preserve or extend. An
engineered structure, perhaps perfect at the time of construction, is perfect for only a period in time. We trust skyscrapers,

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


ships, airplanes, cars and bridges to perform over an expected lifetime. They are designed to have an acceptable, although
non-zero risk level, as they age or when weather or load conditions change. All these lessons must enter into both design and
maintenance to reduce risk. In engineering design, multiple fail-safe principles and redundant systems are included that both
warn of a potential problem and prevent an immediate one. For the oil and gas industry, redundant barriers in well design
perform this purpose with great reliability.

Age versus Era of Construction


From early failures to “old age” wear, time is portrayed to be the enemy of any engineered structure, regardless of the
engineering discipline. Although aging is a significant issue, it must be remembered that failures of the past are what our
knowledge of today is built upon and as learnings progress, the failure rates of a later era will be lower than the era before it.
Everything we know about success is based on mistakes we have made, but only if we learn from them. A key issue with any
operator is how they capture and incorporate learnings into the next design.

For any risk rating, time is a consideration that cannot be ignored. In well construction, time has at least four major influences:

1. Time impacts the knowledge available at the time of well construction. This in turn must reflect the knowledge that
went in to forming the design of the well, the materials available at the time for the construction, and the knowledge-
based regulations that governed construction in that era of time. Failure rates measured in a specific era are artifacts
of that era; they should not be reflective of wells designed and completed today. In well construction, the last 15
years have arguably brought more advances (new pipe alloys, better pipe joints, improved coatings, new cements, and
subsurface diagnostics by seismic and logging delivering better understanding of earth forces) than the previous
fifteen decades of oil and gas operations.

2. Early time failures reflect both quality of well construction and general early component failure (similar to items on a
new car that must be repaired in the first few weeks of operation).

3. Time reflects the potential for natural degradation of materials and changing earth stresses, both natural and man-
made. Structures age; that is inescapable. The impact of aging; however; is highly geographically variable and
controllable to a degree with maintenance. Structures in dry climates and soils often age slowly, while structures in
wet areas, salt spray zones, acid soils and tectonically active areas can be degraded and even destroyed in a few years.
The oldest producing wells, for example, are a century old and many have not leaked, while HPHT, thermal-cycled,
and corrosive environment wells may have a well life of a decade or less before permanent plugging and isolation is
required.

4. Time has also recorded changes in energy source availability from the easily obtainable conventional reservoir
petroleum resources to dependence on and development of resources that are much more difficult to access. This, in
turn has created technology-driven approaches that have been difficult for some, both inside and outside the industry,
to learn and accept.

Potential for downhole leaks to the environment may diminish rapidly as the reservoir pressure is drawn down. Low bottom-
hole pressure wells do not have the driving force to oppose constant hydrostatic pressure of fluids outside the wellbore; hence,
if a leak path is formed through the sequence of barriers, the highest potential is for exterior fluids (usually salt water) to leak
into a wellbore.

One interesting study of 175 wells in Sumatera Island showed the consequences of extending producing past the design time in
an anthropogenic induced environment of insufficient maintenance. This unfortunate combination created a barrier failure rate
of 43% and a well failure rate (with potential exterior leakage) of about 4% of all wells leaking either into or out of the
wellbore (Calosa, 2010).

Older wells that have been maintained through monitoring and repair can be extended past their design life through programs
6 SPE 166142

that assess damage and repair or re-equip wells as needed. An example from the 1932 vintage Bahrain oilfield in the Arabian
Gulf illustrates a moderate degree of casing damage from exterior corrosion, and how the threat was minimized through design
and workovers (Sivakumar, 2004). Casing leaks (one barrier failure), appears to have occurred in many of the 750 wells,
where corrosive salt water flowed into the wells (minimum exterior pollution). In a sequence of design changes dating from
1932 to the early 2000’s, casing damage and barrier failures were driven from 60% frequency to a rare occurrence by
inspection, monitoring and proper maintenance.

In general terms, well construction failures can be caused by inadequate cementing, leaking pipe connections, corrosion, cyclic
loads, thermal extremes, earth stresses, wear/abrasion and other factors. Failures are most commonly reflected by a barrier

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


failure and few of these failures are severe enough to breach all the barriers and pollute ground water. The modern petroleum
industry in the US is nearly one-hundred years old (with a dividing line established by creation of effective cementing
technology and the birth of modern pressure control methods) while US and dedicated worldwide petroleum development
attempts go back over a hundred years before that. Risk reduction has been driven by changes in technology, with an
increasing importance to environmental drivers surfacing in the past 25 years.

From the first U.S. gas wells that used wooden pipe (circa 1820’s) to a few years past entry into the 20th century, zonal
isolation of early wells was haphazard at best. The first true long-term isolation attempts with Portland cement application in
1903 marked the start of the cemented pipe era and the effective two plug cementing system invented by Almond Perkins in
1916 (Lonestar, e-link) moved cement into a proven isolation technique (AnaLog, e-link). Along the way, technology
advances in every field of well development improved the ability to obtain zonal isolation. The major eras of operation and
the notable improvements are shown in Table 2.

Table 2. Approximate Timeline for Pollution Potential by Era.


Time Era Operation Norms Era Potential For Pollution from
Approximation well construction
1820’s to 1916 Cable Tool drilling, no cement isolation, wells openly vented High
to atmosphere.
1916 to 1970 Cementing isolation steadily improving. Moderate
1930’s to Rotary drilling replacing cable tool, pressure control systems Moderate
present & well containment systems developed
1952 to present Hydraulic Fracturing commercialized, reducing number of Low from Fracturing aspects
development wells and required better pipe, couplings and (King, 2012)
cement isolation. (Clark, 1987; Sugden, 2012)
Mid 1960’s to Gas tight couplings and joint make up improving. Moderate for vertical wells, joint
2000 designs improving for horizontal
wells.
Mid 1970’s to Cementing improvements including cement design software, Lower
present data on flow at temperature, dynamic cementing, swelling
cement, flexible, gas-tight and self-healing cements entering
market. (Baumgarte, 1999; Beirute, 1992; , Lockyear, 1990;
Parcevaux, 1984; Ravi, 2002, Chenevert, 1991, Holt; 2012)
1988 to present Multi-frac, horizontal wells, pad drilling reducing Lower
environmental land footprint up to 90%. Improvements in
lower toxicity chemicals from late 1990’s.
2005 to present Well integrity assessments, premium couplings, adding Lower, particularly after 2010
additional barriers and cementing full strings. (Valigura, when state laws were
2005) strengthened on well design.
2008 to present Chemical hazard and endocrine disruptors recognized in Lowest yet, most states caught
fracturing chemicals and sharply reduced. Real time well up with design and inspection
integrity needs being studied to achieve early warning and requirements.
problem avoidance.

Note that the learnings in an era produce lower rates of well failures in the following eras. When comparing failure rates of
wells, regardless of location, well type, or operator; the wells and the problems associated with those wells should be
compared within a single era of well development. Additionally, technology advances have driven improvements in wells,
often without a clear intent to accomplish that goal. For example, rotary drilling enabled development of surface pressure
control systems that eliminated most blowouts, hydraulic fracturing drove better cementing practices and pipe designs, and
horizontal wells reduced the total well count in many areas thereby sharply lowering any risk of surface and subsurface
pollution of fresh water formations. Development of high pressure, high temperature reservoirs drove development of better
SPE 166142 7

seals and pressure handling technology.

Studies of Well Failures


According a review of state-investigated E&P pollution incidences in Ohio (185 cases in ~65,000 wells) and Texas (211 cases
in ~250,000 wells), the majority of pollution incidences were from drilling and completion, production, orphaned wells and
waste disposal, Table 3 and Figure 3. The production well problems were dominated by leaks from pipelines and tanks. These
data include a significant amount of legacy data before the Texas regulations on pits, cementing and barrier design were
changed in 1969 (Kell, 2011). A large overhaul of Texas well regulations in 2012/13 further stiffened well design and

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


implementation requirements.

Many of the Drilling and Completion (D&C) incidents were cement isolation problems, some before the cementing laws were
changed in 1969, 1996 and are being examined again. Fifty-seven of the 75 waste related incidents in Texas during the study
period were legacy issues with disposal pits that were outlawed in 1969. Texas has an industry tax funded program that has
reduced the number of orphaned wells from 18,000 in 2002 to less than 8000 in 2009, and currently is plugging and
abandoning (P&A) 1400 wells per year of the remaining total of orphaned wells (Kell, 2011).

Review of Ohio wells similarly suggested that the pollution incidents were related to early completion practices. Direct quotes
from Kell’s report: “During the 25 year study period (1983-2007), Ohio documented 185 groundwater contamination incidents
caused by historic or regulated oilfield activities. Of those, 144 groundwater contamination incidents were caused by regulated
activities, and 41 incidents resulted from orphaned well leakage. Seventy-six of the incidents caused by regulated activities
(52.7 percent) occurred during the first five years of the study (1983-1987).”

Table 3 - Incidents Investigated & Identified


Pollution Causes (Kell, 2011)
State Ohio Texas
Study Period, years 26 16
# wells producing 65,000 250,000
Num. Cases Investigated 185 211
Site Related 0 0
D&C Related 74 10
Frac Related 0 0
Prod. Related 39 56
Orphan Well Related 41 30
Waste Disposal Related 26 75
P&A Related 5 1
Unknown 0 39

“When viewed in five year increments, the number of incidents caused by regulated activities declined significantly (90.1
percent) during the study period. Seventy-eight percent (113) of all documented regulated activity incidents were caused by
drilling or production phase activities. Improper construction or maintenance of reserve pits was the primary source of
groundwater contamination, which accounted for 43.8 percent of all regulated activity incidents (63) in Ohio.”

“During the 16 year study period (1993-2008), Texas documented 211 groundwater contamination
8 SPE 166142

incidents. More than 35 percent of these incidents (75) resulted from waste management and disposal activities including 57
legacy incidents caused by produced water disposal pits that were banned in 1969 and closed no later than 1984. Releases that
occurred during production phase activities including storage tank or flow line leaks resulted in 26.5 percent of all regulated
activity (TRC) incidents (56) in Texas.”

“During the study period, over 16,000 horizontal shale gas wells, with multi-staged hydraulic fracturing stimulations, were
completed in Texas. Prior to 2008, only one horizontal shale gas well was completed in Ohio. During their respective study
periods, neither the RRC nor the DMRM identified a single groundwater contamination incident resulting from site
preparation, drilling, well construction, completion, hydraulic fracturing stimulation, or production operations at any of these

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


horizontal shale gas wells.” (Comments in quotes from Kell, 2011)

The incident rate of problems from exploration to plug and abandonment, if based on the number of producing wells would be
1 in nearly 1200, while the rate of incidents on total wells drilled in Texas (over 1 million, with over 250,000 in operation),
would be about 1 in 5,000. Note that there were no incidents that directly involved fracturing. This is consistent with recent
studies (King, 2012). These data also indicate that historical environmental incidents associated with oil and gas development
were more commonly associated with above ground fluid handling, leaking tanks or flow lines, use of surface pits to contain
fluids, and less commonly associated with actual well design and construction. Steps have been taken to improve the safety
and protection provided by all aspects of hydrocarbon production and transport, as documented in the data by the reduction in
incidents in recent wells.

On a worldwide scale, the barrier failure rate varies considerably. In a study that looked at well integrity issues, not failures, a
wide variance between barrier failures in regions indicates that conditions and perhaps perceptions may play a role in how well
barrier failures are reported and treated, Table 4 (Feather, 2011).

Table 4 Percentage of wells with single barrier issues in offshore environment.

Geograph. Number Wells Major % wells Source


Area of w/ Problem w/ barrier
Wells barrier issues
issues
Gulf of 14927 6650 Leaks in 45% MMS
Survey,
Mexico Tubulars 2004
North Sea 4700 1600 Tbg Conn., 34% SPE
Forum
UK cement, Survey,
2009
Norway, 2682 482 Tbg Conn. , 18% Norwegian
Petroleum
North Sea cement, Safety
Authority,
2006

Significant reductions in failures can only be accomplished when the causes of the failures as a component and as part of a
system are known and understood. Coupling failures were a high failure area where technology was used to reduce failure
rate. For several decades, pipe connections have been the leading cause of field monitored casing failures, Figure 4.

Pipe thread leaks of some vintages or eras of wells have been estimated to be involved as an element or a direct cause of up to
90% of all barrier failures (Schwind, 2001).
During the 1980’s, problems with wear and brittle failures were reduced by hard banding and alloy changes. Increases in
SPE 166142 9

collapse failure were caused by industry moves to deep water and deeper wells, which increased pipe tension and formation
pressure that leads to increased collapse risk.

In early data (1950 and early 1960’s), exterior pressure testing of over 300,000 casing and tubing couplings made up during
pipe running in the well completion step showed a high number of coupling leaks (1% to almost ~3% of connections) and 50%
and higher frequency of having a leak in a given 8-round threaded pipe string (Kerr, 1965). A number of data gathering and
redesign efforts followed this era and the requirement for better connections in horizontal wells sharply reduced connection
problems.

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


In the early 1960’s (1961 to 1964), an API organized survey of tubular string failures indicated that 86% of the reported casing
string failures and 55% of the reported tubing string failures occurred in connections (Kerr, 1965; Weiner, 1969). These
coupling failures are also indicated in many sustained casing pressure examples.

The API 8-round thread connection is widely popular (used in about 50% of cases) but is not intended for applications above
5000 psi or where gas tight connections are required. The API 8-round thread requires the correct pipe dope (sealant) that both
prevented metal galling during joint makeup and filled the void spaces in the thread pattern to achieve a low pressure gas-tight
seal. Thread make-up procedures and thread dope have been improved and joint make-up measurements are widely used to
reduce leaks (Day, 1990). As late as the 1980’s, tubing leaks were still a prevalent issue with a low time between failures,
Table 5 (Molnes, 1993). When compared to other tubing failures, connection leaks were an obvious problem.

Table 5 Tubing Number MTTF (years)


Failure Mode of Failures
Burst Tubing 3 3113
Collapsed Tubing 7 1334
Restriction in 18 519
Tubing (scale or
collapse)
Tubing parted 6 1557
(broken)
Tubing Leak 182 51
All 216 22
Sintef data: 17 participating companies, 800 total
completion failures & 6600 well yrs operation

In a case history from Norway covering 406 wells, the main problem leading to a barrier failure was tubing (Vignes, 2008),
Table 6, which is consistent with other reports on barrier failure both initially and with increasing age. Looking at the barrier
failures by age shows the principal problems in 5 year age groups up to 19 years, Figure 5, with number of problems per year
for different equipment. Since the well age was not given, only limited deductions can be made on improvements with time for
this data set. The number and distribution of these tubing connection leaks illustrates a technology gap that requires
continuous attention.
10 SPE 166142

Connection reliability focus, beginning in the late 1950’s, has provided improvements in thread design and improvements in
overall barrier reliability. Connection reliability still requires proper make-up: a very human impact. For modern connections,
coupling make-up procedures and metal-to-metal seals in premium threads have eliminated many leak problems from the
1950’s and 1960’s (Payne, 2006). Leaking threads in a single tubular string does not necessarily mean the well will pollute. A
tubing or casing string is a component in a single barrier. An outer barrier can often contain the leaked pressure or fluid until
the leak can be found and repaired (Johns, 2009; Julian, 2007; Cowan, 2007). Cementing the outside of casing, the second
component of the barrier, will eliminate potential for pipe thread leaks if cement covers the thread leak (MacEachern, 2003).

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


Table 6 Percent of 406
Well Integrity Issue Norway (offshore)
wells with the issue
Tubing (connections 7.1%
leak)
Annular Safety Valve 2.2%
(offshore only)
Cement Isolation 2%
Leak
Inner Casing Failure 2%
(connection or
collapse)
Wellhead 1%
Packer 1%
Downhole Safety 0.7%
Valve
Conductor 0.5%
Pack Off 0.5%
Design 0.5%
Gas Lift Valve 0.25%
Fluid Barrier 0.25%
Chemical Injection 0.25%
Valve
Formation 0.25%

Acceptance of premium threads, with superior sealing capacity for critical applications, is helping replace the 8-round thread
in areas requiring gas tight and higher pressure ratings, but connection selection and makeup remains an issue in areas with
older era wells. Premium threaded connections with metal-to-metal have been tested in dozens of North Seal and GOM wells
with exceptionally low incident of leaks or other failures (Dan Gibson, 2000).

Other Barrier Equipment and Importance of Well Maintenance and Monitoring


The failure rate of properly maintained wellheads is low and is commonly limited to seals that isolate the top of tubulars, seals
between the hangers, and valves. Failures in surface barrier systems can be inspected and repaired easily and quickly, without
significant cost or risk. Failure databases on topside (surface) equipment and some subsurface equipment (e.g. subsurface
safety valves) are available through joint industry projects (Sintef, 2010).

Geographic and Well Type Variance of Risk in Well Completions


The type of well and its location is the largest variable on risk to well completions. Wells must be designed to handle the
specific environment, both inside and outside the well. A simplistic view of how certain well types might be viewed for
barrier failure risk is shown in Figure 6.
SPE 166142 11

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


Although the well type comparison of failure estimates is simplistic, it highlights issues created by thermal and pressure
loading in cyclic operations and the effects of corrosive environments inside and outside the well that must be properly
handled in the design phase. Lower risks of horizontal multi-frac wells may be a surprise to some, but Kell’s observations
absence of incidences in the previously referenced Texas study of 16,000 horizontal wells is significant proof and substantiates
engineering expectations that wells properly designed for hydraulic stimulation with modern techniques and materials can
provide excellent environmental protection.

A first pass comparison of failures as they relate to well type is in Table 7 (Moines, 1993). Mean Time to Failure (MTTF) data
are useful as a general comparison of longevity of well equipment as compared to other equipment or similar equipment in
different wells. As recognized in the industry, the produced fluid type has a significant influence on the durability and
longevity of the tubing string. The mean time to failure shows oil production to have the least “wear” impact, and water
injection to have the most wear, since corrosion and thread leaks influenced by corrosion are the dominant failure mechanisms
Table 7. This has significant impact on barrier longevity and reinforces the expectation that wells designed for water injection
typically require more attention during metallurgy selection and ongoing inspection than wells handling a less corrosive
hydrocarbon stream.

Table 7 Tubing Service Time by Well


Type (Sintef Data)
Well Type Service Time Number MTTF
(tbg yrs) of Failures (yrs)
Oil Production 5496 3 3113
Gas/Condensate 1522 7 1334
Production
Water Injection 1756 18 518
Gas Injection 533 6 1556
Data from Molnes, 1993. Total tubing
base of 9333 well-years

Sustained Casing Pressure (SCP)


One of the first signs of a compromised barrier is sustained casing pressure (SCP) or sustained annular pressure (SAP) (API,
2006; Attard, 1991; Bourgoyne, 1999; Soter, 2003, Tinsley, 1980). Sustained casing pressure (SCP) or sustained annular
pressure (SAP) is described as development of a sustained pressure in between the tubing and casing or between a pair of
12 SPE 166142

casing strings that is not caused solely by heating of the well when placed on production. A sustained pressure may be bled-
off quickly but returns over hours or days after the annulus is shut in. Each of the annular spaces (Figure 7) is a separate
pressure vessel and the casing strings are nested to give redundant barriers. Although one barrier may develop a leak,
secondary barriers will contain the pressure and prevent a leak to the outside.

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


Depending on the type, pressure and depth of a well, there may be two to five or more annular spaces in a well. The presence
of sustained pressure in the annulus area indicates two things:

1. There may a seepage-rate flow through at least one tubing string or casing string or through the cement. From this
and other studies, the flow path is most likely either in a pipe coupling in an uncemented steel pipe string or a
microannulus or other type of leak in the cement sheath surrounding the pipe.

2. The pipe and cement on the outside of the annulus is containing the pressure (and fluids) and there may be no
significant leak to the environment outside of the well.

Wells are created with nested redundant barriers that are designed so that a secondary or tertiary barrier backs up the primary
barrier. When sustained annular pressure is detected in a well, the well must be assessed for safety and containment before
further well operations can be considered. If remaining barriers are not capable of containing pressure and fluids, the well
must be shut-in and secured.

Presence of sustained annular pressure appears to be related to geographic location and geology of the sediments (presence of
soft formations and natural seepage paths of gas or oil), the operator’s construction procedures, type of well (corrosive fluid
production, high pressure, cyclic behavior, high temperature formations, etc.), and the age of the well. Complications from
reservoir compaction in young weak rocks often cause casing or tubing collapse requiring changes in design (Bruno, 1992;
Burkowsky, 1981; Fresich, 1996; Hilbert, 1999; Li, 2003).

Barrier and Well Failure Frequency From Case Histories and Databases

Using the extensive presented and published literature from several technical societies, Table 8 captures the ranges of barrier
failures without apparent leaks outside the well and well failure (all barriers in a sequence fail) where fluids may move from
inside the well to the outside (contamination/pollution) or outside to the well to inside the well (intrusion of salt water).
SPE 166142 13

Table 8 Estimates of Barrier or Component Failure & Well Failure (all barriers fail in sequence) for Land Wells
Area # of Type of Wells Type of Risk Barrier Failure Well Failure Data Leaks to Data
Wells Increased by Freq. Range (w/ Range (leak Source GW by Source
barrier failure containment) path sampling

OH 64,830 D&C and shoe D&C – none (test 0.035% in ~0.06% (total) Kell, 2011 Detailed
test fail (74)* fail). (34,000 wells data not
Assume worst. Prod. assumed D&C in 1983- available
Prod. (39) worst in 12 wells 2007), 0.1% (in

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


in 65,000 older wells
worst case.)
TX 253,090 D&C Fail, D&C – none (test 0.02% 0.02% for Kell 2011 0.005% to TGPC,
shoe test fail* fail assumed). older era wells. 0.01% for 2011 &
(10) **Prod Prod – assumed 0.004% for producers TGPC
(56) fail. newer wells &0.03% to 1997
0.07% (inj)
TX 16,000 Horizl Multi- Pollution possible 0*** 0*** Kell 2011 Detailed
frac wells only if well data not
failure is created available
Cedar 671 Vertical Salt creating deep 5.5% UNK Clegg, Detailed
Creek csg collapse 1971 data not
available
Alberta 316,000 Shallow gas, Vent flows and (4.6%) – unk if (4.6%) – unk if Watson, Detailed
tar sand & leaks leaks or single leaks or 2007 data not
muskeg barrier barrier available
* Drilling test, repair required & successful test passage before continuing. ** Includes pipelines
and tanks *** assumed no well integrity loss after reviewing data

Offshore wells are examples of the high end of barrier failure range; although completed production wells in this group present
virtually no potential for groundwater pollution due to absence of subsurface fresh water deposits offshore. The data are
presented for comparison purposes to well construction for land-based wells with higher strength formations.

U.S. Gulf of Mexico (GOM) offshore well data is worst-case barrier failure and pollution potential reporting since the wells
are mainly in gas-charged, often high initial pressure, geologically young and soft marine sands; a combination of geologic
and production extremes that present a difficult well construction environment. Interestingly, for at least the past three decades
in the GOM, excluding the outlier of the Macondo Drilling & Completion blowout in 2010, the amount of oil entering the
GOM from natural seeps is 50 to 500 times the amount spilled or leaked from producing wells (Oil in the Sea, 2003; DOI,
2010), Figure 8.

GOM produced oil leakage by decade as % of oil produced ranges from a low of 0.00005% (1990-99) to 0.007% (1960-69).
14 SPE 166142

Macondo (drilling/completion activity failure) was the clear outlier in GOM oil spills with just less than 1% spillage in 2010.
Natural seeps are shown as constant volume estimates, but many of the natural seeps are episodic in nature.

Sustained Casing Pressure

Industry experience on the Gulf of Mexico Offshore Continental Shelf has shown that the most serious problems with
sustained casing pressure (SCP) have resulted from tubing leaks. Pollution from sustained casing pressure is minor but a risk
of a blowout is increased in some cases (Bourgoyne, 1999 & 2000). Examples of wells with high barrier failure and low well
integrity failure in these offshore environments document the higher barrier failure rate offshore and the positive effect of the

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


multiple barrier system in controlling pollution (reduction of well integrity failures to levels far below barrier failure), Table 9.

Table 9 Sustained Casing Pressure Reports - Estimates of Barrier or Component Failure and Well Failure for Offshore Wells)
Area # Type of Wells Barrier Failure Freq. Range (w/ Integrity Failure Data
Wells containment) Frequency Range (leak Source
in path possible
study
US Gulf 11,498 Platform based 30% overall w/ (tubing-by-prod. csg) 0.01% to 0.05% of Bourgoyne,
of Mexico (3542 wells SCP of 50% of cases. 90% of SCP wells leaked. 0.00005% 1999&2000;
active) have less than 1000 psi. 10% more to 0.0003% of produced Oil in the
Sea, 2003;
serious form of SCP (Wojtanowicz, 2012) oil spilled (1980 - 2009.
DOI, 2010.
US Gulf 4,099 Shoe Test failures 12% to 18% require cement repair to 0 (cement repaired Harris,
of Mexico during drilling* continue drilling before resulting 2008
drilling)
Norway 406 Deviated, 18% 0 Vignes,
offshore 2008
GOM / 2,120 Sand Control 0.5 to 1% 0% subterranean King, 2012
Trinidad ~0.0001% surface via
surface equip erosion
Matagord 17 Compaction 80% to 100% - the high number is Wells routinely shut-in Li, 2003
a Island failures; casing due to high pressure and formation and repaired prior to
623 shear & sand fail compaction. restart.
Sumatera 175 Conventional w/o 43% 1 to 4% Calosa,
maintenance 2010

The sixth example, a study of 175 wells in Sumatera Island showed the consequences of time in a changing environment
coupled with lack of sufficient maintenance. This combination created a barrier failure rate of 43% and a well failure rate
(with potential exterior leakage) of up to 4% (Calosa, 2010).

Why Well Integrity Failures Produce Few Pollution Incidents

As opposed to loss of control during drilling in high pressure reservoirs, such as the Macondo well, actual losses from
completed and producing wells are very low for several reasons:

 Many drilling failures are the result of unexpected high pressure or other drilling-related factors where the pressure
barriers are mostly dynamic (mud weight and BOP control) and before the full range of permanent barriers are
installed that exist in a completed well. The expected frequency of surface releases in production wells (completed
wells) is between one and two orders of magnitude lower during production than during drilling (Fischer, 1996).
Workovers during the life of the producing wells do raise the risk of a release, although the frequency is still about
one order of magnitude lower than during drilling activities (Edmondson, 1996).

 Completed wells are constructed of multiple barriers that have been tested and are monitored where applicable.

 Most importantly, the pressure inside a completed producing oil or gas well drops constantly during primary
production, sharply decreasing the potential from fluids inside the well to flow to the outside the well considering the
outside fluid gradients that increase outside (leak opposing) pressure with increasing depth.

Cementing Basics
Casing and cement barriers in a well result from running casing to an acceptable depth in the formation after drilling a section
of a well. Each section is cased and cemented before the drilling can be continued to deepen the well. If the cement in a section
of the well fails to pass testing requirements, the cement isolation must be repaired before drilling of the next section can
commence. The steel casing provides burst, collapse and tension strength and the cement fill of the annular space between the
SPE 166142 15

drilled hole and the outside of the casing string provides the seal that isolates fluids and pressures. The cement on the outside
of the specific string may come to the surface or to a lower level, depending on the needs of the completion and the threat level
of an isolation failure in the overall design. The amount of cement required to isolate a pipe string depends on the pressures to
be isolated and the quality of the isolation (pressure test and bonding to inner pipe and outer pipe or formation).

The overall effectiveness of the seal depends on the amount of cement fill, the properties of the set cement, elimination of mud
and gas channels within the cement and the bonding of the cement to the pipe and the formation (Watson, 2009; Lockyear
1990).
The cement does not need to be perfect over every foot of the cemented area, but at least some part of the cement column must

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


form a durable and permanent seal that isolates fluids from moving behind the pipe that could contaminate or cause
detrimental mixing of one formation with another. The required amount of this “perfect” cement in a longer column of cement
has been shown to be at least 50 ft, Figure 9. This figure documents twenty-seven short-term comparisons of successful
isolation of pressure in a stacked and sequentially completed pay with as much as 14,000 psi differential between zones in high
pressure gas wells along the soft sediments of the Louisiana Gulf Coast and the Gulf of Mexico. Industry commonly uses
from 200 to 600 ft of cement in overlap sections.

Additional barriers in a well may include a tubing string and packer with a monitored annulus between the tubing (the flow
path) and the inner casing string.

Depending on the area and the specific geologic conditions, there may be one to three barriers in a simple completion with low
risk, and likely two to five barriers in higher risk areas. The differences in design consider pressure, depth, separation,
presence of usable water, corrosion potential and cementing effectiveness in an area, as well as operator expertise in both
design and execution (Smith, 2012; Popa, 2008). Some annuli between casing strings are cemented to surface and some are
left open to enable monitoring of any leak potential from the inside string (O’Brien, 1996). A cement sheath is expected to
provide mechanical support to the casing string and must provide both mechanical and hydraulic isolation for all productive
horizons for the life of the well (Gray, 2009).

Cement Effectiveness in Reducing Pollution Risk


Using results from a study of subsurface water-injection operations in the Williston basin a model was developed to assess the
maximum quantifiable risk that water from water injection wells would reach an underground source of drinking water, The
upper-bound of the probability of injection water escaping the wellbore and reaching a USDW is seven chances in 1 million
well-years where casing and cement cover the drinking-water aquifers. Where surface casings do not cover the USDW’s, the
probability is six chances in one-thousand well years (Michle, 1991). The 1000-to-1 improvement is a testimony to the
efficiency of annular cement.

Cement Bond, Isolation Quality and Bond Testing


The question of how to most effectively test the cement isolation is more involved than most people anticipate. Cement bond
logs, for example, are not always the best investigation tool. Whether or not a bond log should be run depends on the accuracy
and usefulness of the information it gives in a particular application.
16 SPE 166142

Cement bond logs (CBL), which may be sonic or ultrasonic, range from simple averaging instruments similar to those that
came out in 1960 to the more sophisticated models that investigate 60o segments of the cemented annulus around the tool
(Grosmangin, 1996; Flournoy, 1963; Fertl, 1974; Pilkington 1975 & 1992; Bigelow, 1985; Jutten 1991; Goodwin, 1992; Gui,
1996; Frisch 2000 & 2002; Garnier, 2007). A properly run and executed CBL or CET (Cement Evaluation Tool) can provide
some information on cement fill behind the pipe, going past the single area of the cemented shoe and upward and along the
wellbore where a pressure test will not reach.

CBLs are widely used as indicators to help build a good cement job in the first few wells in an area and validate a good
cementing program that will be used on subsequent wells. CBL’s may also be needed after a pressure test has failed on a

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


section of the well during initial construction. For well rework and suspected changes in integrity, many other leak detection
methods offer more accurate determination of leaks than can be produced with a CBL (Johns, 2009; Julian 2007). Cement
bond logs can give a reasonable estimate of bonding and a semi-quantitative idea of presence or absence of larger cement
channels, but will not certify pressure or fluid isolation of a zone. Field performance for a properly run and calibrated CBL is
about 90% in finding channels of 10% or more of total annular space (Albert, 1988). Smaller annular channels are not easily
identifiable with a bond log because of variations in cement composition that create density differences in the cement. This
finding was proven by long-term production without problems, water-free well performance (water isolation) and pressure
measurements over time (Flournoy, 1963). Data in these field tests showed many wells with effective isolation even though
the percentage of acceptable bond ranged from 31% to 75%. Error within the application and interpretation of cement bond
logs has resulted in numerous workovers to repair cement that was not faulty, resulting in high workover costs and a decrease
in the well integrity by unnecessary perforating and attempts to block squeeze cement that was effective in isolating a zone but
“appeared” to have poor bond characteristics on an imprecise CBL.

With the exception of a pressure test, requiring a cement evaluation tool on every cement job appears to be a questionable
policy, with possible detrimental economic and structural consequences for squeeze cementing attempts made on wells with
cement indicated as suspect by CBL investigation, but proven effective by a pressure test and long-term production.

Methane expulsion from the ground may also be associated with well construction by soil and rock disturbance as the well is
being drilled or for a short time after completion (Dusseault, 2000 & 2001). Soil gas disturbance and subsequent venting either
around the casing (short term) or into the well (longer term) is usually found where a large amount of organic materials are
found and either continuously produced or trapped near surface. Examples are swamp-lands, muskegs, Tar sands and some
permafrost areas (Slater, 2010; Macedo, 2012; Dousett, 1997). Of 316,000 wells in Alberta province, most in muskeg areas, an
estimated 4.6% had small surface casing vent flows that lasted a few hours or more after completion. In a 20,500 well subset
of steam injection wells in the shallow tar sands area, almost 15% experienced surface casing vent flows (Nygaard, 2010).
This drives home the point of geographical location and illustrates the difficulty of achieving effective seals in shallow highly
gas productive areas that also are rich in gas seeps and often poorly consolidated.

Natural Surface Seeps of Gas and Oil – Natural Pathways, Natural Pollution Sources
Understanding hydrocarbon movement within natural seeps of oil and gas is critical to the realization that many forms of
methane migration are, in fact, a part of nature. Seep maps from the 1930’s and 1940’s (Link, 1952) show direct correlation
with many fields and hundreds of oil seeps are documented in the Gulf of Mexico and more than a thousand on land in North
America with perhaps as many as 10,000 seeps worldwide (Etiope, 2009). While many uninformed observers may recognize
the correlation between documented hydrocarbon emissions and the presence of oil and gas wells, they may misread the
causality. The presence of natural gas or oil seeps were often the first indicators of highly prospective drilling locations.

Wells are concentrated where significant hydrocarbons accumulations and natural seepage exist, Figure 10; a comment so
evident as to not need stating, except for the strident dialogue against the smallest leakage of hydrocarbon to the surface of the
earth. It is, in fact, these small methane leaks that drew attention to the larger mass of hydrocarbons that lay deep beneath the
surface (ChemTerra). Producible hydrocarbon accumulations have a common trait: natural seeps of gas and oil bring
hydrocarbon fluids to the surface through natural pathways in significant quantities (Wilson, 1974; Kvenvolden, 2003 & 2005; Oil
in the Seas, 2003). Organisms in water and soil have evolved to degrade and use these hydrocarbon sources as food.
SPE 166142 17

Figure 10 – Comparison with Well Density and Natural Seep Location

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


Natural seeps are significant contributors to the global atmospheric methane budget where the worldwide total global methane
seep of 5 to 10 bcf/day ranks just behind the wetlands emissions of 10 to 15 bcf/day (Kvenvolden, 2005; Leifer, 2005; Etiope,
2002; Wilson, 1974).

Table 9 is a compilation of the most severe US oil spills from wells and transport. Since spill volumes are rough estimates,
several sources were used to achieve a reasonable assessment of oil lost. Spills may come from many sources, but
transportation is a primary source. Of the three major spills in the table from the upstream end (exploration and production) of
the oil business, all three are from drilling related causes.

Table 10 Oil spill records for US spills, plus referenced seep and Kuwait spill volume
Rank Year Cause/Source Location Spill Volume Est.
Barrels Gallons Tons
1 1910 Cable Tool Drilling Lakeview California 9,000,000 379,000,000 1,230,000
Gusher (before BOPs developed) onshore
2 2010 Drilling/completion Deep GOM 3,620,000 to 152,000,000 to 492,000 to
Water Horizon (Macondo) 4,500,000 193,000,000 627,000
Seep Every Natural seepage of oil into GOM & 800,000 to 34,000,000 to 110,000 to
Year GOM (non-anthropogenic) CA 1,470,000/yr 62,000,000 /yr 260,000/yr
3 1989 Tanker grounding Alaska 238,000 10,000,000 32,500
4 1976 Tanker grounding Mass. 183,000 7,700,000 25,000
5 2005 Hurricane: primarily pipelines GOM, LA 167,000 7,000,000 23,000
6 1990 Tanker grounding GOM, TX 121,000 5,100,000 17,000
7 1969 Drilling loss of control with no S.B., CA 80,000 to 3,360,000 to 10,900 to
BOP in place 100,000 4,200,000 13,600
Seep Every Natural Oil Seep – Coal Point, California 17,400 to 730,000 to 2,400 to
Year Santa Barbara Channel offshore 26,000 /yr 1,095,000 /yr 3,600 /yr
8 2010 Pipeline corrosion Michigan 20,000 840,000 2,700
9 2000 Tanker grounding GOM, LA 13,500 567,000 1,800
10 2010 Tanker collision GOM, TX 11,000 462,000 1,500
11 2008 Barge collision GOM., LA 10,000 419,000 1,400
Refer 1992 Sabotage – Desert Storm Kuwait - 1,980,000 to 83,000,000 to 270,000 to
ence Onshore 6,000,000 250,000,000 820,000
Sources – Wikipedia Chronologic Spills; NOAA; USGS; Oil In the Seas, Academic press, 2003; Science Daily, NTSB

For comparison, two rows show the yearly ranking of natural seep contribution of oil into waters off California from the Coal
Point seep (largest natural oil seep in the world with daily seep rates of 2,000 to 3,000 gallons of oil and 2 mmscf/d of gas) and
the oil entry by the 600+ natural seeps underlying productive areas of the Gulf of Mexico (GOM). The other reference is the
1992 Kuwait well sabotage where over 600 wells were opened or the wellheads shattered with explosives. Interestingly, the
use of up to 700 lbs. (~320 kg) of high explosives in direct contact with the wellheads only damaged the pipe to a maximum
depth of about 8 to 13 ft (2.4 to 4 m) (Cudd, 1997). The average flow from the GOM and land based natural seeps that are
carried into the GOM every year is estimated to be roughly 25% of the volume spilled by the 99 days of uncontrolled flow
from the Macondo well (Oil in the Seas, 2003).
18 SPE 166142

The information on seeps and spills from tanker and pipeline sources is shown only to put the leak volumes in perspective on a
total oil entry basis, not to attempt to excuse spill behaviors. Seep volumes have a high degree of volume variance since most
seeps are episodic in nature.

Surface flow rates of many oil and gas seeps have dropped sharply in the past century as wells produced fluids from the
reservoirs that fed the seeps (Quigley, 1999). Edwin Drake followed the oil and gas seeps along Oil Creek, near Titusville,
Pennsylvania and in 1859, developed the first purpose-drilled U.S. well for oil that achieved commercial production. The well
hit oil at 69.5 feet, which is shallower than many current Pennsylvania water wells.

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


Hydrocarbons are generated in two ways, by thermal maturation of organic materials that form a range of alkanes from C1
(Methane) to oils of C20+, and by biologic activity that forms most short chain alkanes such as methane gas. Hydrocarbons
and fresh/usable water deposits often share pore space within many formations in their undisturbed states. In many cases, even
fresh water sands such as the Catskill and Lock Haven formations in the northeastern US have sufficient organic material and
burial exposure to create thermogenic gas in the water sands themselves. Several gas fields produced commercial gas from the
Lock Haven and Catskill from 1955 to the late 1960’s, Figure 11, predating all deeper shale gas development in the area
(Molofsky, 2012; Baldassare, 2011).

Figure 11

While methane can be released from activities involving gas well development, properly run scientific studies on large scales
have shown no correlation between gas well development and presence of methane in drinking water (USGS, 2012; Molofsky,
2011; Baldassare, 2011).

Oil also seeps into upper strata through natural pathways. Natural seepage rate of oil is estimated about 600,000 tons per year
or 4 million barrels (170 million gallons) (Wilson, 1974; Kvenvolden, 2003).

Groundwater Variation
Groundwater, brine zones and produced water from oil and gas operations in a specific area do not have a constant
composition. Variance in any groundwater is well documented (Missouri; Nelson, 2012; NETL; Reedy, 2011; Uhlman;
USGS, 2012; Water Encyclopedia; Otten, USGS). Factors in natural variation in ground water and brine zone mineral
composition and methane gas composition include: seasonal variance, barometric pressure changes, changes produced by
runoff, water recharge routes into the reservoir, recharge sources, recharge rates, depth of withdrawal, rate of groundwater
withdrawal and even earth tides. According to groundwater experts, overdrawing or overdrafting a groundwater reservoir
(withdrawing water faster than it can be recharged) can produce notable changes in the reservoir water composition including
pulling contaminants in from above and salt from below (CTIC). This variance makes single point comparisons of water
quality practically worthless. Many fresh groundwater reservoirs are laterally or vertically connected to more saline water
SPE 166142 19

sources. Salinity, within a single reservoir, often varies with depth.

Sudden changes in pressure within a groundwater reservoir will also change the amount of free methane gas by causing gas
breakout into the free gas phase. The only way to assess changes in a groundwater source is to establish a trend range and
include seasonal, withdrawal rate and other variables. Investigations of stray natural gas incidents in Pennsylvania reveal that
incidents of stray gas migration were not caused by hydraulic fracturing of the Marcellus shale (Baldassare, 2011; Molofsky,
2011). The possibility of some gas migration events being related to drilling or cannot be dismissed, since air drilling, when
practiced, may be a cause for temporary upsets in shallow well color and odor, although the worst of these migrations appear
to be in areas with known examples and history of shallow gas flows that predate drilling.

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


Gas migration in the subsurface occurs principally by advective (solubilized in water) transport from areas of high pressure
and is influenced by temporal changes in: barometric pressure, soil/bedrock porosity and permeability, precipitation
influencing pore water levels. Incidents of stray combustible gas are not recent occurrences. Historic gas and oil flows into
water wells and creeks were documented over 200 years ago in New York and Pennsylvania.

Methane in Groundwater and Methane Migration


Methane is the most common gas in groundwater. Shallow methane may be from sources both thermogenic (maturation of
depositional organics in the reservoir) and biogenic (biological breakdown of organic materials carried into the reservoir)
(Burruss, USGS, 2010). Depending on the area of the country and the specifics of the reservoir, groundwater, fresh or saline,
may dissolve and carry methane gas in concentrations of 0 to 28 mg/l. Free (non-dissolved) methane gas exists in many water
reservoirs under the caprock or in rock layers and methane gas is frequently desorbed from organic formations such as coal or
shale as the pressure is reduced by producing the water from these formations. As water is flowed out of a rock formation and
into a wellbore, pressure is lowered, creating opportunity for some dissolved methane gas to come out of the water and
become a free gas phase (similar to the CO2 escaping as a fresh bottle of carbonated soda is opened). Higher draw-downs
(increased differential pressure) will enable more gas to come out of solution. Any free gas will rapidly rise in the well and can
accumulate into the highest part of the well piping system, and, if not vented by proper water well construction, will follow the
moving water when the water tap is opened in the house. This is the cause of burnable gas seen in 200 year old historical
reports on igniting water wells and in more recent dramatic TV and movie shots of burnable gas quantities in faucets and
garden hoses. If the well is constructed correctly, a vent cap allows the methane to escape from the well (Penn, #24) and
prevents most gas accumulations.

Flow paths of methane gas into a water well may be from the fresh water self itself or by communication with old gas wells or
deeper water wells that do not effectively isolate gas charged formations such as coals and thin shale beds, Figure 12.

The frequency of gas appearance in ground water is linked to a number of factors, most of which are far more geographically
or geologically influenced than impacted by oil or gas well presence. Figure 13 shows a number of these factors as positive or
20 SPE 166142

negative influences. Perhaps surprisingly, most of the influences are natural in origin. Wells, particularly those drilled with
air, can play a significant local role in gas migration disturbances (Soeder, 2012). The number of water wells in the US is near
15 million, not counting those that have been dug, driven or drilled and then abandoned. There are very few water well
standards enforced in the US and improperly constructed, poorly maintained and improperly abandoned water wells can be a
primary pathway for aquifer contamination by a variety of materials both from surface and subsurface inflow.

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


Bubbles Around a Wellhead

The presence of bubbles in well cellars or through soil around the wellhead may or may not be an indicator of leaks from the
well. Any disturbance of soil by drilling, digging, pile driving or even walking through a swampy area is frequently
accompanied by release of methane gas; particularly where natural methane in the soil is more highly concentrated (swamps,
muskeg, tundra, etc.). This type of seepage is usually short lived, except in presence of natural seeps fed by deeper reservoirs
along an established seep path. Depending on depth of disturbance, the bubbles may decrease or stop in seconds to days
(Naftz, USGS, 1998). Composition of this gas may be biogenic or thermogenic, since composition of natural seep gas is
commonly from deeper reservoirs therefore overwhelmingly (~96%) thermogenic (Kvendvolden, 2005).

Major Sources of Groundwater Pollution – Where Do Oil and Gas Wells Rank?
Groundwater pollution continues to be a major issue in the US; but what are the primary causes and pathways? The
indisputable major sources of US groundwater contamination over the past few decades are: 1) leaks from underground
storage tanks (gasoline, diesel and chemicals) at filling stations and industrial sites with buried tanks, 2) improper residential
septic systems and 3) poorly constructed landfills (EPA, 2000), Figure 14.
SPE 166142 21

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


In 2000, the first appearance of a potential E&P upstream groundwater pollution source was shallow injection wells at number
17 of the top 20 frequency-reported sources, although other industries also use wells of this type. Oil wells, gas wells and deep
injectors did not make the list.

Case History – Texas Aquifers, Oil & Gas Wells and Pollution
To check the potential of groundwater contamination in a high density oil and gas well environment, data from Texas
Commission on Environmental Quality (TCEQ) and Texas Groundwater Protection Council (TGPC) pollution reports were
reviewed for specific references to reported oil, gas and injection well relevance. The state of Texas was used as an example,
specifically to examine reports of pollution in high density oil and gas well areas on a county-by-county basis. Figure 15 is a
map of the major aquifers of Texas, overlaid with hundreds of thousands of oil and gas wells drilled through those aquifers.
Studies of pollution reports from counties show a higher correlation of oil and saltwater pollution in surface facilities (plants,
compressor stations and tank storage), but few direct downhole results linked to wells (TGPC 1998-2011). Texas is the
number two state in terms of groundwater withdrawal and roughly 80% of the withdrawals are used for agriculture and
municipal water supply. Aquifers of varying quality and quantity underlay between 80 to 90% of Texas lands. Reports of
water quality from properly built water supply wells affecting public health or crops is absent from the literature, indicating
either very little impact of gas and oil wells on the majority of aquifers, or a curious and unheard-of lack of reporting.

Figure 15 Aquifers, Oil and Gas Wells in Texas


22 SPE 166142

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


Texas pollution records were examined for major causes of pollution and for possible links to oil and gas wells. The reported
information (volumes of pollutant released was not available), Figure 16, showed strong but decreasing report frequency of
impact from underground gasoline and diesel storage tanks at filling stations, and variable exposure to manufacturing
chemicals, solvents, waste oils, and very small incidents of spills involving road transport of oil.

Figure 16 Reported Incidents of Groundwater Pollution in Texas (TGPC records, 2000 to 2011)

The percentage of pollution reports that TCEQ and/or TGPC identified as under Texas Railroad Commission (TRC) authority
varied between less than 1% to a maximum of about 10% of the total new pollution reports each year. Further analysis of
these reports showed a breakdown of cases where surface facilities (tanks, separators, gas plants, compressors and gathering
lines), and pipelines were about 90% of the pollution incidents where TRC had jurisdiction. The remaining reports of
pollution, roughly less than 1% of the total reported each year, concerned the 200,000 to 250,000 wells that are producing,
injecting or shut-in in Texas during the time period. All leaks were ascribed to wells whether they had been investigated or not
– a worst-case approach. No leaks from abandoned wells were listed although these may have been addressed in the Texas
Orphan well program which ranks and then P&A’s over 1400 wells per year with proceeds from a tax on permits and
operators. The breakdown is shown in Figure 17.
SPE 166142 23

Figure 17 Specifics of TGPC and TRC pollution reporting from 2000 to 2011.

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


Conclusions
1. Current redundant barrier well design with nested cemented casing strings are effective in sharply reducing pollution
potential from oil and gas wells.
2. One or more barriers in a properly designed and constructed oil or gas well may fail without creating a pollution
pathway or significantly raising the risk of groundwater pollution.
3. The overall risk of groundwater pollution from a producing well is extremely low for most quality operations.
4. Individual barrier failure rates and well failure rates vary widely with type of well, geographical location area and
maintenance culture of the operator.
5. Individual well barrier element failure rates are often one to two orders of magnitude higher than well integrity
failures where all barriers in a protection sequence fail and pollution can or does happen.
6. Failure rates of well barriers and well integrity failures, measured on wells constructed in a specific time era are
artifacts of that era; they are not identical to failure rates of wells designed and completed today.
7. Leak numbers that could be justifiably allocated to oil, gas or injection wells indicate an overall percent of leaking
wells in the range of 0.03% to 0.005% of wells in service at the time of the report; however; areas with older wells
and surface facilities tend to dominate the complaints while newer development areas such as the Barnett have fewer
confirmed pollution incidents (lower % of wells involved with problems). Accuracy of numbers for frequency of
pollution reporting by wells in this type of a survey is obviously dependent on the ability to detect a leak and the
ability to identify the source once it is reported.
8. The formation fluid pressure differential between exterior formations and the well environment favors fluid flow into
the well instead of out of the well because downhole pressure in an oil or gas wells is reduced as reservoir fluid is
removed.
9. Gas migration potential varies primarily with geography and appears to be highest where natural seeps of gas or oil
are present. Gas from discrete natural seeps is overwhelmingly thermogenic gas. Large volumes of biogenic gas area
released from swamps, landfill, tundra, etc.
10. Methane gas migration from deep drilling does not appear to be connected to oil and gas production, but other factors
of drilling and development, such as unknown location of improperly abandoned wells may be factors.
11. Improperly constructed and maintained water wells and improperly plugged oil or gas wells may be conduits for
methane migration into fresh water supplies.
12. Groundwater composition may change with numerous factors, most of which are unrelated to oil and gas operations.

Acknowledgements
The authors thank the management of Apache and Willis Consulting Group for permission to publish.

Technical content reviewers include Randy Valencia (Apache), Mike Vinson (Consultant), Dean Wehunt (Chevron), Dan
Daulton (Baker Hughes), Mike Bahorich (Apache), Mark Zoback (Stanford University), Stephen Tipton (Newfield).
24 SPE 166142

References
1. Aadnoy, B.S., Belayneh, M., Arriado, M., Flateboe, R.: “Design of Well Barriers To Combat Circulation Losses,” SPE
Drilling & Completions, Vol. 23, No. 3, September 2008.
2. ALARP: http://www.hse.gov.uk/risk/theory/alarp.htm
3. ALARP and De Minis Risk: http://www.stb07.com/technical-safety/alarp.html
4. AnaLog: http://www.logwell.com/tech/cbl/cement_problems.html
5. Albert, L.E., et.al.: “A Comparison of CBL, RBT and PET Logs in a Test Well with Induced Channels,” JPT (Sept, 1988)
1211-16.
6. Anders, J., Rossberg, S., Dube, A., Engel, H., Andrews, D.: “Well Integrity Operations at Prudhoe Bay, Alaska,” SPE

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


Production and Operations, May 2008, p280-286.
7. API RP90, 2006, Annular Casing Pressure Management for Offshore Wells, Washington D.C., API.
8. API TR 10TR2, Shrinkage and Expansion in Oilwell Cements, first edition, 1997, Washington D.C.: API.
9. Attard, M.: “The Occurrence of Annular Pressures in the Northwest Hutton Field: Problems and Solutions,” SPE 23136, SPE
Offshore Europe, Aberdeen, 3-6 September 1991.
10. Baldassare, F.: “The Origin of some natural gases in Permian through Devonian Age Systems in the Appalachian Basin & the
relationship to incidents of stray gas migration,” Prepared for EPA Technical Workshops for Hydraulic Fracturing Study,
February 24, 2011. http://www.epa.gov/hfstudy/theoriginofsomenaturalgasesinpermianthroughdevonianagesystems.pdf
11. Battelle: Joint Industry Research Program, Subsurface Hydrocarbon Containment: Challenges and R&D Needs, October
2012.
12. Baumgarte, C., Thierelin, M., Klaus, D.: “Case Studies of Expanding Cement to Prevent Microannular Formation,” SPE
56535, SPE Annual Meeting and Exhibition, 3-6 October 1999. Houston, TX, USA.
13. Bazzari, J.A.: “Well Casing Leaks History and Corrosion Monitoring Study, Wafra Field,” SPE 17930, SPE Middle East Oil
Show, 11-14 March 1989, Bahrain.
14. Beirute, R.M., Wilson, M.A., Sabins, F.L.: “Attenuation of Casing Cemented with Conventional and Expanding Cements
Across Heavy-Oil and Sandstone Formations,” SPE Drilling Engineering, Sept 1992, 2000.
15. Bigelow, E.L.: “A Practical Approach to the Interpretation of Cement Bond Logs,” SPE 13342, JPT, 1985.
16. Bonn, R.J.C.: “Operational Lessons from the UKCS Hydrocarbon Leaks Database, SPE 46641, 1998 SPE International
Conference on Health, Safety and Environment, Caracas, Venezuela, 7-10 June 1998.
17. Bosma, M., Ravi, K., van Driel, W., Schreppers, G.J.: “Approach to Sealant Selection for the Life of the Well,” SPE 56536,
SPE Annual Meeting and Exhibition, 3-6 October 1999. Houston, TX, USA.
18. Bourgoyne, A. Jr., Scott, S.L., Regg, J.B.: “Sustained Casing Pressure in Offshore Producing Wells,” OTC 11029, Offshore
Technology Conference, Houston, TX, 3-6 May 1999.
19. Bourgoyne Jr., A.T.J., S.L. Scott, and W. Manowski, A Review of Sustained Casing Pressure(SCP) Occurring on the OCS,
2000, Final Report Submiteted to MMS.
20. Browning, L.A., Smith, J.D.: “Analysis of the Rate and Reasons for Injection Well Mechanical Integrity Test Failure,” SPE
25967, SPE/EPA Exploration and Production Environmental Conference, 7-10 March 1993, San Antonio, TX, USA.
21. Bruno, M.: “Subsidence Induced Well Failure,” SPE Drilling Engineering, June 1992, Vol. 7, No. 2, June 1992.
22. Burkowsky, M, Ott, H., Schillinger, H.: “Cement Pipe-In-Pipe Casing Strings Solved Field Problems, “World Oil, October
1981, pp. 143-47.
23. Burruss, R.C., Laughrey, C.D.: “Covariation of carbon and hydrogen isotopic compositions in natural gas: separating
biogenic, thermogenic, and abiotic (inorganic CO2 reduction) sources,” USGS presentation.
http://pa.water.usgs.gov/projects/energy/stray_gas/presentations/2_830_Burruss.pdf
24. Calosa, W.J., Sadarta, B., Ronaldi: “Well Integrity Issues in Malacca Strait Contract Area,” SPE 129083, SPE Oil and Gas
India Conference and Exhibition, Mubai, India, 20-22 Jan 2010.
25. Carter, L.G.: “Effect of Bore-Hole Stresses on Set Cement,” MS Thesis, U. of Oklahoma, Norman, OK, (1968).
26. CDC: “water-related Diseases and Contaminants in Private Wells,”
http://www.cdc.gov/healthywater/drinking/private/wells/diseases.html
27. ChemTerra International: “Introduction to Seeps”. http://www.chemterra.com/seep-introduction/
28. Chenevert, M.E., Shrestha, B.K.: “Chemical Shrinkage Properties of Oilfield Cements, SPE Drilling Engineering, Vol 6,
No.1. March, 1991, pp 37-43.
29. Clark, H.C.: “Mechanical Design Considerations for Fracture-Treating Down Casing Strings,” SPE Drilling Engineering,
June 1987.
30. Clegg, J.D.: “Casing Failure Study – Cedar Creek Anticline,” SPE Journal of Petroleum Technology, Vol. 23, No. 6, June
1971.
31. Corneliussen, K., Sorli, F., Brandanger Haga, H., Tenold, E., Menezes, C., Grimbert, B., Owren, K.: “Well Integrity
Management System (WIMS) – A Systematic Way of Describing the Actual and Historic Integrity Status of Operational
Wells, SPE 110347, SPE Annual Technical Conference and Exhibition, Anaheim, CA, USA. 11-14 November 2007.
32. Cowan, M.: “Field Study Results Improve Squeeze Cementing Success,” SPE 106765, SPE Production Operations
Symposium, 31 March – 3 April 2007, Oklahoma City, OK, USA.
33. Crow, W.: “Studies on Wellbore Integrity,” Proceedings of the 2nd Wellbore Integrity Network Meeting, Princeton, New
Jersey, 28-29 March 2006.
34. CTIC: “Groundwater & Surface Water: Understanding the Interaction, 2nd edition,” Purdue.
http://www.ctic.purdue.edu/media/files/Ground%20Water%20and%20Surface%20Water.pdf
35. Cudd, Bobby Joe: personal conversation on his on-site experience with dozens of well control operations on burning wells
and explosives damaged wells following Desert Storm, 1992.
36. D’Alesio, P., Poloni, R., Valente, P, Magarini, P.A.: “Well-Integrity Assessment and Assurance: The Operational Approach
SPE 166142 25

for Three CO2 Storage Fields in Italy,” SPE Production & Operations, May 2011, pp 140-148.
37. Day,J.B., Moyer, M.C., Hirshberg, A.J.: “New Makeup for API Connections,” SPE Drilling, Vol. 5, No. 3, September 1990,
pp 233-238.
38. Demong, K., Hands, R., Affleck, B.: “Advanvements in Efficiency, in Horn River Shale Stimulation,” SPE 140654, SPE
Hydraulic Fracturing Technology Conference, 24-26 January, 2011, The Woodlands TX USA.
39. DOI - Department of the Interior: “Increased Measures for Energy Development on the Outer Continental Shelf,” May 27,
2010.
40. Dethlefs, J.C.: “Assessing Well Integrity Risk: A Qualitative Model,” SPE 142854, SPE Drilling and Completions, Vol. 27,
No. 2., pp294-302, June 2012.

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


41. Dousett, J., Akit, S., MacDonald, R.: “Reduction of Surface Casing Vent in East-Central Alberta Using Improved Drilling
and Cementing Techniques,” Petroleum Society of Canada 97-185, Petroleum Conference of the South Saskatchewan
Section, Oct 19-22, 1997.
42. Dugay, A., Polo, J., Pages, B., Belgaroui, J.: “Zero Tolerance to Hydrocarbon Leaks: A Multidisciplinary Success,” SPE
154743, SPE Mid East Health, Safety Security and Environment Conference and Exhibition, Abu Dhabi, UAE, 2-4 April
2012.
43. Dusseault, M.B., Bruno, M.S., Barrera, J.: “Casing Shear: Causes, Cases, Cures,” SPE Drilling & Completions, Vol. 16, No.
2, June 2001, pp98-107.
44. Dusseault, M.B., Grau, M.N., Nawrocki,, P.A.: “Why Oilwells Leak: Cement Behavior and Long Term Consequences,” SPE
64733, International Oil and Gas Conference and Exhibition, Beijing, China, 7-10 November 2000.
45. Edmondson, J.N., Hide, D.K.: “Risk Assessment of Hydrocarbon Releases During Workover and Wireline Operations on
Completed Wells on Offshore Platforms,” SPE 36913, 1996 SPE European Petroleum Conference, 22-24 October, 1996,
Milan, Italy.
46. Etiope, G.: “A Global Dataset of Onshore Gas and Oil Seeps: A New Tool For Hydrocarbon Exploration,” Istituto Nazionale
di Geofisica e Vulcanologia, Roma, etiope@ingv.it. Oil and Gas Business, 2009.
47. Etiope, G. (2009): “Natural Emissions of methane from geological seepage in Europe,” Environment, 43, 1430-1443, doi:
10.1016/j.atmosenv. 2008.03.014, 2009.
48. Etiope, G., Ciccioli, P. (2009): “Earth’s degassing – A missing ethane and propane source. Science, 323, 5913, 478, doi:
10.1126/science. 1165904.
49. Etiope, G., Klusman, R.W.: “Geologic Emissions of Methane to the Atmosphere,” 2002, Chemisphere 49, 777-789.
50. EPA, 2000 National Water Quality Inventory, Chapter 6, Ground Water, Figure 6-5, p52.
51. EPA, Coalbed Methane Extraction: Detailed Study Report, December 2010, EPA-820-R-10-022.
52. Faul, R., Kelm, C., Slocum, T., Crook, R.: “Fine-Grind Cement Aids, GOM Plug-and-Abandon Operations,” OTC, 3-6 May
1999, Houston, TX, USA.
53. Fertl, W. H., Pilkington, P. E., Scott, J.B.:”A Look at Cement Bond Logs,” Journal of Petroleum Technology, Vol. 26, No. 6,
June, 1974, pp 607-617.
54. Feather, K.: “Better Well Integrity,” US-Norway Technology Partnership Conference, Minimizing Oil Spills & Discharges to
Sea, Houston, TX, USA, 30 March 2011.
55. Fischer, F.: “Identification and Quantification of Failure Rate to Effectively Manage Sour Gas Production Systems,” 1996
SPE HSE in OIL and Gas E&P Conference, 9-12 June 1996, New Orleans, LA, USA.
56. Fleming, K.N., Silday, F.A.: “A risk informed defence-in-depth framework for existing and advanced reactors, Reliability
Engineering & System Safety, 78(3), 205-225.
57. Flournoy, R.M., Feaster, J.H.: “Field Observations on the Use of the Cement Bond Log and Its Application to the Evaluation
of Cementing Problems,” SPE 632, Society of Petroleum Engineers, New Orleans, LA, October 3-6, 1963.
58. Fresich, J.T., Arguello, J.G., Wawersik, W.R., Deitrick, G.L., de Rouffignac, E.P., Myer, L.R., Bruno, M.S.: “Three-
Dimensinal Geomechanical Simulation of Reservor Compaction and Implications for Well Failure in the Bellridge
Diatomite,” SPE Annual Technical Conference and Exhibition, 6-9 October 1996. Denver, CO, USA.
59. Frisch, G. O’Mahoney, L., Mandal, B.: “Examination of Cement and Casing Evaluation Logs,” IADC/SPE 77212, Jakarta,
Indonesia, 9-11 Sept, 2002.
60. Frisch, G., Graham, Griffith, J.: “A Novel and Economic Processing Technique Using Conventional Bond Logs and
Ultrasonic Tools for Enhanced Cement Evaluation,” SPWLA 41th, Annual Logging Symposium, June 4-7, 2000.
61. Garnier, A., Fraboulet, B., Saint-Marc, J., Bois, A.-P.: Characterization of Cement Systems to Ensure Cement Sheath
Integrity,” OTC, 18754, Houston, TX, USA, 30 April – 3 May 2007.
62. Gibson, Dan: personal communication on Dan’s experience on Amoco and BP North Sea Operations Testing jointed
connections, 2000.
63. Goodwin, K.J., Crook, R.J.: “Cement Sheath Stress Failure,” SPE 20453, SPE Drilling Engineering, Vol. 7, No. 4.,
December, 1992, pp 291-296.
64. Gray, K.E., Podnos, E., Becker, E.: “Finite-Element Studies of Near-Wellbore Region During Cementing Operations, Part 1,”
SPE Drilling and Completion, Vol 24, No. 1, March 2009.
65. Grosmangin, M., Kokesh, P.P., Majani, P.: “A Sonic Method for Analyzing the Quality of Cementation of Borehole
Casings,” SPE Journal of Petroleum Technology. Vol. 13, No. 2, pp 165-171. Feb, 1961.
66. Gui, H., Summers, T. D., Cocking, D.A., Greaves, C.: “Zonal Isolation and Evaluation for Cemented Horizontal Liners,” SPE
Drilling and Completion, December, 1996.
67. Haddon, W.J.: The basic strategies for reducing damage from hazards of all kinds,” Hazard Prevention, Sept-Oct, 1980, pp. 8-
12.
68. Hale, A., Gossens, L., Ale, B, Bellamy, L., Post, J., Oh, J., et.al. (2004). Manading safety barriers and controls at the
workplace,” PSAM 7-ESREL 2004, Berlin.
26 SPE 166142

69. Handal, A.: “Safety Barrier Analysis and Hazard Identification of Blowout Using Managed Pressure Drilling Compared with
Conventional Drilling,” 2013 IADC/SPE Managed Pressure Drilling and Underbalanced Operations Conference &
Underbalanced Operations Conference & Exhibition,April 17-18, 2013, San Antonio, TX, USA.
70. Harris, K., Grayson, G., Langlinais, J.: “Obtaining Successful Shoe Tests in the Gulf of Mexico: Critical Cementing Factors,”
SPE 71388, SPE Annual Technical Conference and Exhibition, 30 Sept. – 3 Oct 2001, New Orleans, LA.
71. Hilbert, L.B., Gwinn, R.L., Moroney, T.A., Deitrick, G.L.: “Field Scale and Wellbore Modeling of Compaction-Induced
Casing Failures,” SPE Drilling & Completions, Vol. 14, No. 2, June 1999.
72. Holt, C., Lahoti, N.: “Dynamic Cementation Improves Wellbore Construction and Reduces the Hazards of Groundwater
Contamination in Shale Plays, SPE 162739, SPE Canada Unconventional Resources Conference, 30 Oct – 1 November 2012,

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


Calgary, Alberta, Canada.
73. Hostettler, F., Kvenvolden, K.: “Studies of Contaminated Ground Water Yield New Insights into Degradation of Diesel
Fuel,” USGS Soundwaves Monthly Newsletter, August 2002.
74. Humphreys, A., Ross, R.: “Overcoming the Loss of a Primary Barrier in an HPHT Well – Investigation and Solution,” SPE
105736, SPE/IADC Drilling Conference, 20-22 February 2007, Amsterdam, The Netherlands.
75. Johns, J.E., Blount, C.G., Dethlefs, J.C., Loveland, M.J., McConnell, M.L., Schwartz, G.L. Julian, J.J.: “Applied Ultrasonic
Technology in Wellbore-Leak Detection and Case Histories in Alaska North Slope Wells,” SPE Production Operations,
Volume 24, No. 2, May 2009, pp 225-232.
76. Julian, J.J., King, G.E., Johns, J.E., Sack, J.K., Robertson, D.B.: “Detecting Ultrasmall Leaks with Ultrasonic Leak Detection,
Case Histories from the North Slope, Alaska,” SPE 108906, International Oil Conference and Exhibition in Mexico, 27-30
June 2007, Veracruz, Mexico.
77. Jutten, J., Toma, I., Morel, Y., Ferreol, B.: “Integration of Cement Job Data for Better Bond Index Interpretation”, SPE
21690, OKC, April 7-9, 1991.
78. Kappel, W., Hamilton, H.: “Dissolved Methane Found in Some New York Groundwater,” US DOI, USGS, Released
9/4/2012, http://www.usgs.gov/newsroom/article.asp?ID=3391#.UFIa5FHGx8F .
79. Kell, S.: “State Oil and Gas Agency Groundwater Investigations and Their Role in Advancing Regulatory Reforms, A Two
State Review: Ohio and Texas,” August 2011, Ground Water Protection Council, 2011.
80. Kerr, H.P.: “Thread Leaks in Tubing and Casing Strings,” API 65-014, Drilling and Production Practice, 1965.
81. Khalifeh, M., Saasen, A., Hodne, H., Vralstad, T.: “Techniques and Materials for North Sea Plug and Abandonment
Operations,” OTC 23915, 6-9 May 2013, Houston, TX, USA.
82. King, G.E.: “Hydraulic Fracturing 101: What Every Representative, Environmentalist, Regulator, Reporter, Investor,
University Professor, Neighbor and Engineer Should Know About Estimating Frac Risk and Improving Frac Performance in
Unconventional Gas and Oil Wells,” SPE 152596, SPE Hydraulic Fracturing Technology Conference, The Woodlands, TX,
USA, 6-8 February 2012.
83. King, G.E.: “Thirty Years of Shale Gas Fracturing: What Have We Learned?” SPE 133456, 19-22 September, 2010,
Florence, Italy.
84. King, G.E., Wildt, P.J., O’Connell, E.: “Sand Control Completion Reliability and Failure Rate Comparison with a Multi-
Thousand Well Database,” SPE 84262, SPE Annual Technical Conference and Exhibition, 5-8 Oct. 2003, Denver, CO, USA.
85. Kvenvolden, K.A., Cooper, C.K.: “Natural Seepage of Crude Oil into the Marine Environment,” Geo-Mar Lett, (2003) 23:
140-146.
86. Kvenvolden, K.A., Rogers, B.W.: “Gaia’s Breath – global methane exhalations,” Marine and Petroleum Geology 22 (2005)
579-590.
87. Li, X., Mitchum, F.L., BrunoM., Pattillo. P.D., Willson. S.M.: “Compaction, Subsidence, and Associated Casing Damage and
Well Failure Assessment for the Gulf of Mexico Shelf Matagorda Island 623 Field,” SPE 84553, SPE Annual Technical
Conference and Exhibition, 5-8 October 2003, Denver, CO, USA.
88. Liang, Q., J.: “Casing Thermal Stress and Wellhead Growth Behaviors Analysis,” SPE 157977, SPE Asia Pacific Oil and Gas
Conference, 22-24 October 2012, Perth Australia.
89. Leifer, L., Luyendyk, B.P., Boles, J., Clark, J.R.: “Natural marine seepage blowout: Contribution of Atmospheric methane,”
Global Biogeochemical Cycles, Vol. 20, GB308, doi: 10.1029/2005GB002668
90. Link, W.K.: “Significance of Oil and Gas Seeps in World Oil Exploration,” Bulletin of the American Association of
Petroleum Geologists, Volume 36, Number 8, August 1952.
91. Lockyear, C.F.: “Cement Channeling: How to Predict and Prevent,” SPE Drilling Engineering, Vol. 5, No. 3, September
1990, pp 201-208.
92. Lonestar Securities; http://lonestarsecurities.com/Book-CH-III.htm
93. MacEachern, D.P., Algu, D.R., Cowan, M., Harris, K., Snell, E.: “Advances in Tieback Cementing,” SPE/IADC Drilling
Conference, 19-21 February 2003, Amsterdam, The Netherlands.
94. Macedo, K.G., Schneider, J.W., Sylvestre, C.J., Masoor, Q.: “Elimination of Surface Casing Vent Flow and Gas Migration in
the Lloydminster Area,” SPE 157922, SPE Heavy Oil Conference, Calgary, Alberta, 12-14 June 2012.
95. Mansouri F.A., Alam, M.A., “Sources of Hydrocarbon Leaks and Spills in Upstream Oil Industries, Its Potential, Reasons
And Preventative Measures, SPE 111725, 2008 SPE International Conference on Health, Safety and Environment, Nice
France, 15-17 April 2008.
96. MSC - Marcellus Shale Coalition: “Methane: An Element of Nature,” http://marcelluscoalition.org/2013/03/methane-an-
element-of-nature/ March 4, 2013.
97. MSC - Marcellus Shale Coalition: “Recommended Practices: Pre-Drill Water Supply Problems,” MSC RP 2012-3 Auguest
28, 2012, Updated March 19, 2013.
98. MSC – Marcellus Shale Coalition: “Recommended Practices: Responding to Stray Gas Incidents,” MSC RP 2012-4 October
16, 2012.
SPE 166142 27

99. Martland, R., Mann, P. D.: “Examining The Sustainability of E&P Major Accident Prevention Design Principles in A
Changing Global Environment and Comparisons With the Rail, Nuclear and Aviation Industries,” SPE 156910, SPE/APPEA
International Conference on Health, Safety, and Environment in Oil and Gas Exploration and Production, 11-13 September,
2012 Perth, Austrlia.
100. Michle, T.W., Koch, C.A.: “Evaluation of Injection Well Risk Management in the Williston Basin,” SPE Journal of
Petroleum Technology, June 1991, pp 737-741.
101. Michie & Associates Inc., “Status Update of Williston Basin Class II Injection Well Study,” UIPC/EPA – Region VI Class I
& II Injection Well Symposium, Dallas, TX – May 9, 1989.
102. Michie & Associates Inc., “Evaluation of Injection Well Risk Management Potential In the Williston Basin,” Underground

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


Injection Practices Council Research Foundation, September 1989.
103. Missouri Department of Natural Resources: “Why Groundwater Levels Change,”
http://www.dnr.mo.gov/env/wrc/docs/whywaterlevelschange.pdf
104. MIT: The Facts on Fracking Methane Emissions, The Energy Collective, http://theenergycollective.com/mark-
green/149821/mit-facts-fracking-methane-emissions
105. Molnes, E., Sundet, I.: “Reliability of Well Completion Equipment,” SPE 26721, SPE Offshore European Conference, 7-10
September 1993, Aberdeen Scotland.
106. Molofsky, L.J., Connor, J.A., Farhat, S. K., Wylie, A.S.Jr., Wagner, T.: “Methane in Pennsylvania water wells unrelated to
Marcellus shale fracturing,” Oil and Gas Journal, Dec 5, 2011, pp 54-67.
107. Naftz, D.L., Hadley, H.K., USGS: “Determination of Methane Concentrations in Shallow Ground Water and Soil Gas near
Price, Utah., January 1998
108. Nelson, D.: “Natural Variations in Composition of Groundwater,” Presented at Groundwater Foundation Annual Meeting,
November 2002. (Drinking Water Program, Oregon Department of Human Services, Springfield Oregon).
109. NETL, “A Guide to Practical Management of Produced Water from Inshore Oil and Gas Operations in the United States,”
110. NPC: “Plugging and Abandonment of Oil and Gas Wells,” Paper 2-25, Prepared by the Technology Subgroup of the
Operations & Environmental Task Group. Sep 5, 2011.
http://downloadcenter.connectlive.com/events/npc091511/Prudent_Development-091511.pdf
111. Nygaard, R.: “Well Design and Well Integrity, Wabamun Area CO2 Sequestration Project (WASP),” Institute for Sustainable
Energy, Environment and Economy (ISEEE), University of Calgary, January 4, 2010.
112. Oakley, J.S.: “Using Accident Theories to Prevent Accidents,” 05-534, ASSE Professional Development Conference and
Exposition, June 12-15, 2005, New Orleans, LA, USA.
113. O’Brien, T.B.: “A Case Against Cementing Casing-Casing Annuli,” SPE 35106, IADC/SPE Drilling Conference, 12-15
March 1996, New Orleans, LA, USA.
114. Oil in the Sea III, Committee on Oil in the Sea, National Research Council of the National Academies, The National
Academic Press, Washington DC, 2003.
115. Otten, J.K., Mercier, T.: “Produced Water Brine and Stream Salinity,” USGS, http://water.usgs.gov/orh/nrwww/Otten.pdf
116. Parcevaux, P.A., Sault, P.H.: “Cement Shrinkage and Elasticity: A New Approach to a Good Zonal Isolation,” SPE 13176,
SPE Annual Technical Conference and Exhibition, 16-19 September, 1984, Houston, TX, USA.
117. Payne, M.L., Schwind, B.E., Park, J., Coker, O., Wu, J., Postler, D., Comet, A.: “Project aimes to qualify tubular
connections,” Drilling Contractor, May/June 2006, p60-61.
118. Penn State Extension: Methane Gas and Its Removal from Water Wells in Pennsylvania,” Water Facts # 24,
119. Pilkington, P.E., Fertl, W.H.: “Field Tests of Cement Bond Logging Tools,” The Log Analyst, Vol. XVI, No. 4. July-Aug,
1975.
120. Pilkington, P.E.”Cement Evaluation – Past, Present and Future,” Journal of Petroleum Technology, Vol 44, No. 2, Feb 1992.
Pp 132-140.
121. Popa, A., Popa, C., Malamma, M., Hicks, J.: “Case-Based Reasoning Approach for Well Failure Diagnostics and Planning,”
SPE 114229, SPE Western Regional and Pacific Section AAPG Joing Meeting held in Bakersfield, CA, 31 March – 2 April,
2008.
122. PSA (2005). “Trend in risk levels – summary report,” Phase 5 (2004). Stvanger, Norway, The Petroleum Safety Authority.
123. Quigley, D.C., Hornafius, J.S., Luyendyk, B.P., Francis, R.D., Clark, J., Washburn, L.: “Decrease in natural marine
hydrocarbon seepage near Coal Point, California, associated with offshore oil production,” Geology, November 1999, v. 27,
no. 11, p. 1047-1050.
124. Randohl, P., Carlsen, I.M.: “Assessment of Sustained Well Integrity on the Norwegian Continental Shelf,” Fourth Mtg of
IEA-GHG Wellbore Integrity Network, 18-19 March 2008, Paris, France.
125. Ravi, K., Bosma, M., Gastebled, O.: “Improve the Economics of Oil and Gas Wells by Reducing the Risk of Cement
Failure,” IADC/SPE Drilling Conference 26-28 Feb. 2002, Dallas, TX, USA.
126. Ravi, K., Bosma, M., Gastebled, O.: “Safe and Economic Gas Wells Through Cement Design for Life of the Well,” SPE
75700, SPE Gas Technology Symposium, 30 April – 2 May 2002, Calgary, Alberta, Canada.
127. Reedy, R.C., Scanlon, B.R., Walden, S., Strassberg, G.: “Naturally Occurring Groundwater Contamination in Texas,” Bureau
of Economic Geology, The University of Texas at Austin, October, 2011, Contract Number 1004831125.
128. Ririe, G.T., Sweeney, R.E.: “Comparison of Hydrocarbon Gases in Solis From Natural Seeps and Anthropogenic Sources,”
Northwest Gas Association, http://info.ngwa.org/gwol/pdf/930159740.PDF
129. Ritchie, N.W.: “Journey to Zero: Aspiration Versus Reality,” SPE 164963, SPE European HSE Conference and Exhibition,
London, UK, 16-18 April 2013.
130. Rao, Vikram: Shale Gas, The Promise and the Peril, RTI Press, 2012
131. Schwind, B.E., Payne, M.L., Otten, G.K., Pattillo, P.D.: “Development of Leak Resistance in Industry Standard OCTG
Connections Using Finite Element Analysis and Full Scale Testing,” SPE 13050, Offshore Technology Conference, 30 April
28 SPE 166142

– 3 May 2001, Houston, TX, USA.


132. Schwind, B.E., Curington D.W., Miller, R.A.: “LRFD Derived Performance of a Qualified API Connection Population,” OTC
7490, 27th Annual, OTC, Houston, TX. 1-4 May 1995.
133. Siegel, D.: “Reductionist hydrology: ten fundamental principles,” Hydrol. Process. 22, 4967-4970 (2008).
134. Siegel, D.: “Errors in Myers’ Marcellus Shale Groundwater Paper from Start to Finish,” Energy in Depth, 13 May 2013.
http://eidmarcellus.org/marcellus-shale/errors-in-myers-marcellus-shale-groundwater-paper-from-start-to-finish/8761/
135. Sintef: PDS Data Handbook, 2010 Edition
136. Sivakumar, V.C., Janahi, A.,: “Salvage of Casing Leak Wells on Artificial Lift in a Mature Oil Field,” SPE 88749, Abu Dhabi
International Conference and Exhibition, 10-13 October 2004, Abu Dhabi, UAE.

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022


137. Shahri, M.A., Schubert, J.J., Amani, M.: “Detecting and Modeling Cement Failure in High-Pressure/High-Temperature
(HP/HT) Wells, Using Finite Element Method (FEM),” International Petroleum Technology Conference, 21-23 Nov. 2005,
Doha, Qatar.
138. Sklet, S.: “Safety barriers: Definition, Classification, and performance,” Journal of Loss Prevention, 2006.
139. Sklet, S.: “Hydrocarbon releases on oil and gas production platforms: Release scenarios and safety barriers,” Journal of Loss
Prevention, 2006.
140. Slater, H.J.: “The Recommended Practice for Surface Casing Vent Flow and Gas Migration,” SPE 134257, SPE Annual
Technical Conference and Exhibition, Florence Italy, 19022 September, 2010.
141. Smith, L., Milanovic, D., Lee, C.H., Billingham, D.,: “Modeling and Predicting of the Corrosion of Onshore Well Casings,”
NACE Corrosion 2012, 11-15 March, 2012, Salt Lake City, UT.
142. Soeder, D.J.: “Field Test of an Alternative Hypothesis for Stray Gas Migration from Shale Gas Development,” NETL,
Morgantown, WV, Presentation for GWPC Stray Gas Forum, July 25, 2012 Cleveland, Ohio
143. Soter, K, Medine, F., Wojtanowicz, A.K.: Improved Techniques to Alleviate Sustained Casing Pressure in a Mature Gulf of
Mexico Field,” SPE 84556, SPE Annual Conference and Exhibition, 5-8 Denver, 2003, Denver CO, USA.
144. Steinsvik, D., Vegge, A.: “Multiple Barrier Thinking – A Road To A Proactive Culture,” SPE 111607, SPE International
Conference on Health, Safety and Environment in Oil and Gas Exploration and Production, 15-17 April, Nice, France.
145. Sueker, J.K., Clark, B.L., Cramer, G.H.: “Isotope Techniques for Discriminating Methane Sources,” 8th International
Petroleum & Biofuels Environmental Conference, Houston, TX, November 8-10, 2011.
146. Sugden, C., Johnson, J., Chambers, M., Ring, G, Suryanarayana, P.V.: “Special Considerations in the Design Optimization of
the Production Casing in High Rate, Multistagge-Fractured Shale Wells,” SPE Drilling and Completions, December 2012,
pp459-472.
147. Tewari, R.D., et.al.: “The Importance of Well Construction and Well Integrity for Reservoir Management – A Mature Field
Experience in Sundan,” SPE 100813, SPE Asia Pacific Oil & Gas Conference and Exhibition, Adelaide, Australia, 11-13
September, 2006.
148. TGPC, Texas Groundwater Protection Committee: “Joint Groundwater Monitoring and Contamination Report – 2011,” SFR-
056/11, August 2012.
149. TGPC, Texas Groundwater Protection Committee: “Joint Groundwater Monitoring and Contamination Report – 1997,” SFR-
56/98, June 1998.
150. Tinsley, J.M., Miller, E.C., Sabins, F.L., Sutton, D.L.: “Study of Factors Causing Annular Gas Flow Following Primary
Cementing,” SPE JPT, August 1980, pp 1427-1437.
151. Turner, M, et.al.: “Permeable Reactive Barriers: Lessons Learned/New Directions,” February 2005, Prepared by Interstate
Technology & Regulatory Council – Permeable Reactive Barriers Team.
152. U.S. Government, Injection Well Regulations, 40 CFR 146, http://www.gpo.gov/fdsys/pkg/CFR-2012-title40-
vol24/xml/CFR-2012-title40-vol24-part146.xml
153. USGS: “Coal-Bed Methane: Potential and Concerns,” http://pubs.usgs.gov/fs/fs123-00/fs123-00.pdf
154. USGS: “Magnetic Surveys for Locating Abandoned Wells,” http://pubs.usgs.gov/fs/fs-0163-95/FS163-95.html
155. USGS: “Effects of Ground-Water Development on Ground-Water Flow To and From Surface Water Bodies.”
156. USGS: “Dissolved Methane in New York Groundwater,” http://pubs.usgs.gov/of/2012/1162/
157. Uhlman, K., Rock, C., Artiola, J.: “Arizona Drinking Water Well Contaminants,” University of Arizona Cooperative
Extension
158. Vaziri, H., Allam, R., Kidd, G., Bennett, C., Grose, T., Robinson, P., Malyn, J.: “Sanding: A Rigerous Examination of the
Interplay Between Drawdown, Depletion, Start-up Frequency and Water Cut,” SPE 89895, 2006, SPE Journal of Petroleum
Technology, Vol. 21, No. 4. 430-440.
159. Valigura, G.A., Tallin, A.: “Connections for HPHT Well Applications and Connection Leak Probability,” SPE HPHT Sour
Well Design Applied Technology Workshop, 17-19 May 2005, The Woodlands, TX. USA.
160. Veil, J. , Puder, M.G., Elcock, D., Redwick, R.J., Jr.: “A White Paper Describing Produced Water from Production of Crude
Oil, Natural Gas, and Coal Bed Methane,” Prepared for US DOE, NETL, Contract W-31-109-Eng-38, January 2004.
161. Vignes, B., Aadnoy, B.S.: “Well-Integrity Issues Offshore Norway,” SPE112535, IADC/SPE Drilling Conference, 4-6 March
2008, Orlando, FL., USA.
162. Vignes, B.: “Qualification of Well Barrier Elements – Long Term Test, Test Medium and Temperatures,” SPE 138465, SPE
European Health, Safety and Environmental Conference in Oil and Gas Exploration in Oil and Gas Exploration and
Production, 22-24 February, 2011, Vienna, Austria.
163. Water Encyclopedia: Natural Composition of Fresh Water. http://www.waterencyclopedia.com/En-Ge/Fresh-Water-Natural-
Composition-of.html
164. Watson, T.L., Bachu, S.: “Evaluation of the Potential for Gas and CO2 Leakage Along Wellbores,” SPE Drilling and
Completion, March 2009.
165. Weiner, P.D., True, M.E.: “A Method of Obtaining Leakproof API Threaded Connections in High-Pressure Gas Service, API
SPE 166142 29

69-040, API Drilling and Production Practice, 1969.


166. Wilson, R.D., Monaghan, P.H., Osanik, A., Rogers, M.A.: “Natural Marine Oil Seepage,” Science, May 24, 1974.
167. Wojtanowicz, A.K., Nishikawa, S., Rong, X.: “Diagnosis and Remediation of Sustained Casing Pressure in Wells,” US DOE,
Baton Rouge, July 2001.
168. Woodyard, A.H.: “Risk Analysis of Well Completion Systems,” SPE Journal of Petroleum Technology, Vol. 34., No. 4, April
1982.
169. Wu, J., Knauss, M., Kritzler, T.: “Casing Failures in Cyclic Steam Injection Wells,” SPE 114231, IADC/SPE Asia Pacific
Drilling Technology Conference and Symposium, Jakarta, Indonesia, 25-27 August, 2008, Jakarta, Indonesia.
170. Zinkham, R.E., Goodwin, R.J.: “Burst Resistance of Pipe Cemented into the Earth,” JPT, September, 1962, pp1033-40.

Downloaded from http://onepetro.org/SPEATCE/proceedings-pdf/13ATCE/3-13ATCE/D031S044R001/1531975/spe-166142-ms.pdf by Saudi Aramco, Baraket Mehri on 25 February 2022

You might also like