You are on page 1of 49

In-Situ

Combustion
Enhanced Oil Recovery

Prep by: Taha Gebril


Table of Contents

PREP BY: TAHA GEBRIL ...............................................................................................................

TABLE OF CONTENTS ................................................................................................................ I

LIST OF FIGURES ...................................................................................................................... IV

LIST OF TABLES ....................................................................................................................... IV

CHAPTER 1 ................................................................................................................................... 1

1.1 OIL PRODUCTION ............................................................................................................ 1

1.2 OIL PRODUCTION SCHEMES ............................................................................................. 1

1.2.1 PRIMARY RECOVERY ......................................................................................................................1


1.2.2 SECONDARY RECOVERY ..................................................................................................................1
1.2.3 TERTIARY RECOVERY ......................................................................................................................2

1.3 ENHANCED OIL RECOVERY (EOR) .................................................................................. 2

1.3.1 TECHNIQUES ...................................................................................................................... 2

1.3.1.1 GAS INJECTION ..........................................................................................................................3


1.3.1.2 THERMAL METHODS ...................................................................................................................3
1.3.1.3 CHEMICAL INJECTION ..................................................................................................................4
1.3.1.4 MICROBIAL INJECTION.................................................................................................................4

1.4 ECONOMIC COST AND BENEFIT OF EOR ....................................................................... 5

1.5 CONCEPT OF THERMAL EOR ............................................................................................ 5

1.6 HISTORY OF THERMAL EOR ............................................................................................. 6

CHAPTER 2 ................................................................................................................................. 10

IN-SITU COMBUSTION ............................................................................................................ 10

2.1 CONCEPT OF ISC ................................................................................................................ 10

2.2 ASSETS AND LIABILITIES OF IN-SITU COMBUSTION PROCESS ............................. 11

2.3 ASSETS OF IN-SITU COMBUSTION PROCESS .............................................................. 11

2.4 LABORATORY STUDIES ................................................................................................... 12

2.4.1 OXIDATION CELLS ........................................................................................................................12


I
2.4.2 COMBUSTION TUBE STUDIES .........................................................................................................13

2.5 QUALITATIVE DESCRIPTION OF IN-SITU COMBUSTION ......................................... 14

2.6 MECHANISM UNDER PROCESSES OF IN-SITU COMBUSTION WORKING ............ 15

2.6.1 FORWARD DRY COMBUSTION ........................................................................................................16


2.6.2 FORWARD WET COMBUSTION .......................................................................................................16
2.6.3 COMPARISON OF WET AND DRY COMBUSTION...................................................................................17
2.6.4 REVERSE COMBUSTION ...............................................................................................................19

2.7 OPERATING PRACTICES ................................................................................................... 21

2.7.1. COMPRESSORS ..........................................................................................................................21


2.7.2 AIR REQUIREMENT ......................................................................................................................22
2.7.3 IGNITION ...................................................................................................................................22
2.7.4 WELL DESIGN AND COMPLETIONS ..................................................................................................22
2.7.5 INJECTION AND PRODUCTION PRACTICES .........................................................................................23

2.8 PILOT TEST ........................................................................................................................ 26

2.9 PATTERNS SWEEP, INVASION AND DISPLACEMENT EFFICIENCIES .................. 26

2.9.1 SWEEP EFFICIENCIES ....................................................................................................................28

2.10 OIL CONSUMED IN-SITU .............................................................................................. 29

2.11 OIL RECOVERY BY ISC ................................................................................................. 29

2.12 FACTORS AFFECTING IN-SITU COMBUSTION ........................................................ 30

2.12.1 LATERAL AND VERTICAL EXTENT OF RESERVOIRS..............................................................................30


2.12.2 VERTICAL DEPTH .......................................................................................................................31
2.12.3 RESERVOIR THICKNESS ...............................................................................................................31
2.12.4 STRUCTURAL ATTITUDE AND DIP ..................................................................................................32
2.12.5 OVERBURDEN COMPETENCE........................................................................................................32
2.12.6 RESERVOIR HETEROGENEITIES ......................................................................................................32
2.12.7 ROCK PROPERTIES .....................................................................................................................33
2.12.8 SAND UNIFORMITY AND TEXTURE.................................................................................................33
2.12.9 PERMEABILITY ..........................................................................................................................34
2.12.10 POROSITY ..............................................................................................................................34
2.12.11 OIL SATURATION .....................................................................................................................34
2.12.12 COMPOSITION OF RESERVOIR MATRIX.........................................................................................34
2.12.13 EFFECT OF WELL SPACING.........................................................................................................35

2.13 INJECTION OF OXYGEN-ENRICHED AIR OR PURE OXYGEN .............................. 35


II
2.14 LIMITATIONS OF COMBUSTION PROCESS ................................................................ 36

2.15 DESIGN CONSIDERATIONS ............................................................................................ 36

2.16 USE OF IN-SITU COMBUSTION ..................................................................................... 37

2.16.1 USE AS A PRIMARY RECOVERY PROCESS ..........................................................................................37


2.16.2 USE AS A SECONDARY RECOVERY PROCESS ......................................................................................37
2.16.3 USE AS A TERTIARY RECOVERY PROCESS..........................................................................................38

2.17 ECONOMIC EVALUATION .............................................................................................. 38

2.18 DISADVANTAGES OF IN-SITU COMBUSTION ........................................................... 39

2.19 CURRENT STATUS OF IN-SITU COMBUSTION .......................................................... 39

2.20 CASE STUDY ..................................................................................................................... 40

2.20.1 RESERVOIR FLUID PROPERTY .......................................................................................................40


2.20.2 IMPLEMENTATION SCHEME .........................................................................................................40
2.20.3 PROCESS MONITORING ..............................................................................................................41
2.20.4 AIR INJECTION PRESSURE ............................................................................................................43
2.20.5 PRODUCTION PERFORMANCE ......................................................................................................43
2.21 SUMMARY – CONCLUSION .............................................................................................................43

REFERENCE ............................................................................................................................... 44

III
List of Figures

Figure 1.1. Numbers of EOR projects in USA


Figure 1.2. Production of EOR methods in USA
Figure 1.3. Oil viscosity as a function of temperature and gravity
Figure 1.4. Typical heavy oil Viscosity vs Temperature
Figure 2.1. In situ combustion drive mechanism
Figure 2.2. Differential thermal analysis of a crude oil (From Burger and Sahuquet,1972)
Schematic representation of in situ combustion process and the various zones as
Figure 2.3. formed in the oil reservoir
Schematic illustration of the in-situ combustion process. (After Nelson and
Figure 2.4. McNeil 1961; courtesy of the Oil and Gas Journal.)
Temperature distribution in the reservoir during a fireflood. (After Nelson and
Figure 2.5. McNeil, 1961; courtesy of the Oil and Gas Journal.)
Figure 2.6. Temperature profiles in dry and wet combustion
Figure 2.7. Comparison of wet and dry combustion
Figure 2.8. Temperature profile in reverse combustion
Schematic illustration of the reverse combustion process. (After Berry and
Figure 2.9. Parrish. 1960; courtesy of the Society of Petroleum Engineers of AIME.)
Variation of ignition time as a function of temperature for several air-injection
Figure 2.10. pressures. (After Dietz and Weijdema, 1968; courtesy of the Producers Monthly.)
Figure 2.11. Typical production well design
Figure 2.12. Water needed to cool hot wells
Figure 2.13. (a) Open inverted and (b) confined direct five-spot
Figure 2.14. Unburned zones, pattern sweep, and invasion efficiencies

List of Tables

Table 1.1. Categories of heavy oil


Table 2.1. Recovery Efficiency of In-Situ Combustion Compared to Other EOR Methods
Table 2.2. Application of Pollution Control Systems to a Fireflood Project
Table 2.3. Pattern Sweep Efficiency

IV
Chapter 1

1.1 Oil Production

Is the process by which usable oil is drawn out from the beneath the earth’s surface. Geologists and
geophysicists use seismic surveys to search for geological structures that may form oil reservoirs. The "classic"
method includes making an underground explosion nearby and observing the seismic response, which provides
information about the geological structures underground. However, "passive" methods that extract
information from naturally occurring seismic waves are also used.
Other instruments such as gravimeters and magnetometers are also used in the search for petroleum.
Extracting crude oil normally starts with drilling wells into an underground reservoir. When an oil well has been
tapped, a geologist (known on the rig as the "mudlogger") will note its presence.
Historically in the United States, in some oil fields the oil rose naturally to the surface, but most of these fields
have long since been used up, except in parts of Alaska. Often many wells (called multilateral wells) are drilled
into the same reservoir, to an economically viable extraction rate. Some wells (secondary wells) may
pump water, steam, acids or various gas mixtures into the reservoir to raise or maintain the reservoir pressure
and economically extraction rate. The oil well is created by drilling a long hole into the earth with an oil rig. A
steel pipe (casing) is placed in the hole, to provide structural integrity to the newly drilled well bore. Holes are
then made in the base of the well to enable oil to pass into the bore. Finally a collection of valves called a
"Christmas Tree" is fitted to the top; the valves regulate pressures and control flow.

1.2 Oil Production Schemes

1.2.1 Primary Recovery

During the primary recovery stage (also called the 'Kareem Stage'), reservoir drive comes from a number of
natural mechanisms. These include: natural water displacing oil downward into the well, expansion of the
natural gas at the top of the reservoir, expansion of gas initially dissolved in the crude oil, and gravity drainage
resulting from the movement of oil within the reservoir from the upper to the lower parts where the wells are
located. Recovery factor during the primary recovery stage is typically 5-15%.
While the underground pressure in the oil reservoir is sufficient to force the oil to the surface, all that is
necessary is to place a complex arrangement of valves (the Christmas tree) on the well head to connect the
well to a pipeline network for storage and processing. Sometimes pumps, such as beam pumps and electrical
submersible pumps (ESPs), are used to bring the oil to the surface; these are known as artificial lifting
mechanisms.

1.2.2 Secondary Recovery

Over the lifetime of a well, the pressure falls. At some point there is insufficient underground pressure to force
the oil to the surface. After natural reservoir drive diminishes, secondary recovery methods are applied. These
rely on supplying external energy to the reservoir by injecting fluids to increase reservoir pressure, hence
increasing or replacing the natural reservoir drive with an artificial drive. Secondary recovery techniques
increase the reservoir's pressure by water injection, natural gas reinjection and gas lift, which
injects air, carbon dioxide or some other gas into the bottom of an active well, reducing the overall density of
fluid in the wellbore. The typical recovery factor from water-flood operations is about 30%, depending on the
properties of the oil and the characteristics of the reservoir rock. On average, the recovery factor after primary
1
and secondary oil recovery operations is between 35 and 45%.

1.2.3 Tertiary Recovery

Enhanced, or Tertiary oil recovery methods, increase the mobility of the oil in order to increase extraction.
Thermally enhanced oil recovery methods (TEOR) are tertiary recovery techniques that heat the oil, reducing
its viscosity and making it easier to extract. Steam injection is the most common form of TEOR, and it is often
done with a cogeneration plant. This type of cogeneration plant uses a gas turbine to generate electricity, and
the waste heat is used to produce steam, which is then injected into the reservoir. This form of recovery is used
extensively to increase oil extraction in the San Joaquin Valley, which yields a very heavy oil, yet accounts for
ten percent of the United States' oil extraction.[citation needed] Fire flooding (In-situ burning) is another form
of TEOR, but instead of steam, some of the oil is burned to heat the surrounding oil.
Occasionally, surfactants (detergents) are injected to alter the surface tension between the water and the oil
in the reservoir, mobilizing oil which would otherwise remain in the reservoir as residual oil.
Another method to reduce viscosity is carbon dioxide flooding.
Tertiary recovery allows another 5% to 15% of the reservoir's oil to be recovered. In some California heavy
oil fields, steam injection has doubled or even tripled the oil reserves and ultimate oil recovery. For example,
see Midway-Sunset Oil Field, California's largest oilfield.
Tertiary recovery begins when secondary oil recovery is not enough to continue adequate extraction, but only
when the oil can still be extracted profitably. This depends on the costof the extraction method and the
current price of crude oil. When prices are high, previously unprofitable wells are brought back into use, and
when they are low, extraction is curtailed.
The use of microbial treatments is another tertiary recovery method. Special blends of the microbes are used
to treat and break down the hydrocarbon chain in oil, making the oil easy to recover. It is also more economical
versus other conventional methods. In some states such as Texas, there are tax incentives for using these
microbes in what is called a secondary tertiary recovery. Very few companies supply these, however, companies
like Bio Tech, Inc. have proven very successful in waterfloods across Texas.

The amount of oil that is recoverable is determined by a number of factors, including the permeability of the
rock, the strength of natural drives (the gas present, pressure from adjacent water or gravity), porosity of the
reservoir rock, i.e the rock storage capacity, and the viscosity of the oil. When the reservoir rocks are "tight", as
in shale, oil generally cannot flow through, but when they are permeable, as in sandstone, oil flows freely.

1.3 Enhanced Oil Recovery (EOR)

Is the implementation of various techniques for increasing the amount of crude oil that can be extracted from
an oil field. Enhanced oil recovery is also called tertiary recovery (as opposed to primary and secondary
recovery). According to the US Department of Energy, there are three primary techniques for EOR: thermal
recovery, gas injection, and chemical injection. Sometimes the term quaternary recovery is used to refer to
more advanced, speculative, EOR techniques. Using EOR, 30 to 60 percent, or more, of the reservoir's original
oil can be extracted, compared with 20 to 40 percent using primary and secondary recovery.

1.3.1 Techniques

There are three primary techniques of EOR: gas injection, thermal injection, and chemical injection. Gas
injection, which uses gases such as natural gas, nitrogen, or carbon dioxide (CO2), accounts for nearly 60

2
percent of EOR production in the United States. Thermal injection, which involves the introduction of heat,
accounts for 40 percent of EOR production in the United States, with most of it occurring in California. Chemical
injection, which can involve the use of long-chained molecules called polymers to increase the effectiveness of
waterfloods, accounts for about one percent of EOR production in the United States. In 2013, a technique called
Plasma-Pulse technology was introduced into the United States from Russia. This technique can result in
another 50 percent of improvement in existing well production.

1.3.1.1 Gas Injection

Gas injection or miscible flooding is presently the most-commonly used approach in enhanced oil recovery.
Miscible flooding is a general term for injection processes that introduce miscible gases into the reservoir. A
miscible displacement process maintains reservoir pressure and improves oil displacement because the
interfacial tension between oil and water is reduced. This refers to removing the interface between the two
interacting fluids. This allows for total displacement efficiency. Gases used include CO2, natural gas or nitrogen.
The fluid most commonly used for miscible displacement is carbon dioxide because it reduces the
oil viscosity and is less expensive than liquefied petroleum gas. Oil displacement by carbon dioxide
injection relies on the phase behaviour of the mixtures of that gas and the crude, which are strongly dependent
on reservoir temperature, pressure and crude oil composition.

1.3.1.2 Thermal Methods

In this approach, various methods are used to heat the crude oil in the formation to reduce its viscosity and/or
vaporize part of the oil and thus decrease the mobility ratio. The increased heat reduces the surface tension
and increases the permeability of the oil. The heated oil may also vaporize and then condense forming improved
oil. Methods include cyclic steam injection, steam flooding and combustion. These methods improve the sweep
efficiency and the displacement efficiency. Steam injection has been used commercially since the 1960s in
California fields. In 2011 solar thermal enhanced oil recovery projects were started in California and Oman, this
method is similar to thermal EOR but uses a solar array to produce the steam.
- Steam flooding: is one means of introducing heat to the reservoir by pumping steam into the well with a
pattern similar to that of water injection. Eventually the steam condenses to hot water; in the steam zone
the oil evaporates, and in the hot water zone the oil expands. As a result, the oil expands, the viscosity
drops, and the permeability increases. To ensure success the process has to be cyclical. This is the principal
enhanced oil recovery program in use today.

- In-Situ Combustion (Fire flooding): it works best when the oil saturation and porosity are high. Combustion
generates the heat within the reservoir itself. Continuous injection of air or other gas mixture with high
oxygen content will maintain the flame front. As the fire burns, it moves through the reservoir toward
production wells. Heat from the fire reduces oil viscosity and helps vaporize reservoir water to steam. The
steam, hot water, combustion gas and a bank of distilled solvent all act to drive oil in front of the fire
toward production wells. There are three methods of combustion: Dry forward, reverse and wet
combustion. Dry forward uses an igniter to set fire to the oil. As the fire progresses the oil is pushed away
from the fire toward the producing well. In reverse the air injection and the ignition occur from opposite
directions. In wet combustion water is injected just behind the front and turned into steam by the hot
rock. This quenches the fire and spreads the heat more evenly.

3
1.3.1.3 Chemical Injection

The injection of various chemicals, usually as dilute solutions, have been used to aid mobility and the reduction
in surface tension. Injection of alkaline or caustic solutions into reservoirs with oil that have organic
acids naturally occurring in the oil will result in the production of soap that may lower the interfacial
tension enough to increase production. Injection of a dilute solution of a water-soluble polymer to increase
the viscosity of the injected water can increase the amount of oil recovered in some formations. Dilute solutions
of surfactants such as petroleum sulfonates or biosurfactants such as rhamnolipids may be injected to lower
the interfacial tension or capillary pressure that impedes oil droplets from moving through a reservoir. Special
formulations of oil, water and surfactant, microemulsions, can be particularly effective in this. Application of
these methods is usually limited by the cost of the chemicals and their adsorption and loss onto the rock of the
oil containing formation. In all of these methods the chemicals are injected into several wells and the
production occurs in other nearby wells.

- Polymer Flooding: Polymer flooding consists in mixing long chain polymer molecules with the injected
water in order to increase the water viscosity. This method improves the vertical and areal sweep
efficiency as a consequence of improving the water/oil Mobility ratio.
- Surfactants may be used in conjunction with polymers; they decrease the surface tension between the oil
and water. This reduces the residual oil saturation and improves the macroscopic efficiency of the process.
Primary surfactants usually have co-surfactants, activity boosters, and co-solvents added to them to
improve stability of the formulation.
- Caustic flooding is the addition of sodium hydroxide to injection water. It does this by lowering the surface
tension, reversing the rock wettability, emulsification of the oil, mobilization of the oil and helps in
drawing the oil out of the rock.

1.3.1.4 Microbial Injection

Microbial injection is part of microbial enhanced oil recovery and is rarely used because of its higher cost and
because the developments is not widely accepted. These microbesfunction either by partially digesting
long hydrocarbon molecules, by generating biosurfactants, or by emitting carbon dioxide (which then functions
as described in Gas injection above).
Three approaches have been used to achieve microbial injection. In the first approach, bacterial cultures mixed
with a food source (a carbohydrate such as molasses is commonly used) are injected into the oil field. In the
second approach, used since 1985, nutrients are injected into the ground to nurture existing microbial bodies;
these nutrients cause the bacteria to increase production of the natural surfactants they normally use to
metabolize crude oil underground. After the injected nutrients are consumed, the microbes go into near-
shutdown mode, their exteriors become hydrophilic, and they migrate to the oil-water interface area, where
they cause oil droplets to form from the larger oil mass, making the droplets more likely to migrate to the
wellhead. This approach has been used in oilfields near the Four Corners and in the Beverly Hills Oil
Field in Beverly Hills, California.
The third approach is used to address the problem of paraffin wax components of the crude oil, which tend to
precipitate as the crude flows to the surface, since the Earth's surface is considerably cooler than the petroleum
deposits (a temperature drop of 9-10-14 °C per thousand feet of depth is usual).

4
1.4 Economic Cost and Benefit of EOR

Adding oil recovery methods adds to the cost of oil —in the case of CO2 typically between 0.5-8.0 US$ per
tonne of CO2. The increased extraction of oil on the other hand, is an economic benefit with the revenue
depending on prevailing oil prices. Onshore EOR has paid in the range of a net 10-16 US$ per tonne of
CO2 injected for oil prices of 15-20 US$/barrel. Prevailing prices depend on many factors but can determine the
economic suitability of any procedure, with more procedures and more expensive procedures being
economically viable at higher prices. Example: With oil prices at around 90 US$/barrel, the economic benefit is
about 70 US$ per tonne CO2. The U.S. Department of Energyestimates that 20 billion tons of captured
CO2 could produce 67 billion barrels of economically recoverable oil.
It is believed that the use of captured, anthropogenic carbon dioxide, derived from the exploitation
of lignite coal reserves, to drive electric power generation and support EOR from existing and future oil and gas
wells offers a multifaceted solution to U.S. energy, environmental, and economic challenges. There is no doubt
that coal and oil resources are finite. The U.S. is in a strong position to leverage such traditional energy sources
to supply future power needs while other sources are being explored and developed. For the coal industry,
CO2 EOR creates a market for coal gasification byproducts and reduces the costs associated with carbon
sequestration and storage.

1.5 Concept of Thermal EOR

Thermal Enhanced Oil Recovery (TEOR) normally used in heavy oil reservoirs. Heat is propagated into the
reservoir to raise the temperature which lowers the oil viscosity and makes it easier to flow and decreases the
interfacial tension between oil and injected fluid. Heat energy moves to the formation, convection in the
reservoir fluids and conduction through the formation rock. If the oil viscosity is high, the flow rate in the
reservoir will be correspondingly low and it will take a long time to effectively deplete the reservoir. Oil viscosity
can be significantly reduced by raising the temperature of the oil in the reservoir. Unfortunately, it is not
possible to raise the temperature of the oil without raising the temperature of water in the pore spaces of the
rock and of the reservoir rock itself. Often a considerable amount of energy will have to be added to the
reservoir to accomplish this objective. There are two basic techniques for heating the reservoir:

1. Injecting steam or hot gases into the reservoir


2. Injecting either air or oxygen into the reservoir and burning some of the oil in the reservoir .This technique
is called in situ combustion

Dolberry Oil estimates that steam accounts for 52% of the market methods utilized for EOR. In comparison with
gas injection, carbon dioxide is 31% and nitrogen is 17%. Steam provides additional pressure that produces greater
oil production as the additional heat assists in losing the crude oil in the "pay zone" surrounding the well.

5
1.6 History of Thermal EOR

Thermal Enhanced Oil Recovery technique appeared in the fifties of the nineteenth century ( 1950s ) and spread
widely in many countries, especially the United States and was the first way appearance and widely used and
continued to spread while Gas EOR appearance in seventies of the nineteenth century ( 1970s ) and Chemistry
EOR in eighties o0f the nineteenth century ( 1980s ), Thermal recovery methods, even though their growth rate
may have peaked in the United States, will continue growing elsewhere in the world, where massive heavy oil and
tar sands are found and where environmental restrictions may not be as severe as those in the United States. As
the world’s conventional oil production starts to decline within 30 years, attention will be focused on the almost
immeasurable quantities of heavy oil and tar sands, much of which can only be produced by thermal methods,
which represents approximately 45% of worldwide oil reserves.

Five periods of evolution for Thermal Enhanced Oil Recovery (TEOR):

1. Early and mid-fifties when mainly laboratory work was carried out
2. Late fifties and early sixties when the processes were tested in field pilot projects.
3. Mid-sixties to early seventies when large steam-soak projects were started and research experiments were
carried out in large physical models.
4. Mid-seventies to early eighties when the number of steam projects has been increasing fast. The design of
these projects is being carried out with the aid of numerical simulators.
5. The present, when new techniques and applications for thermal methods are under investigation.

The following figures showing EOR history in the United States, which is considered as one of the world's most
commonly used of EOR and TEOR.

Figure 1.1. Numbers of EOR projects in USA

6
Figure 1.2. Production of EOR methods in USA

Heavy crude oil or extra heavy crude oil is any type of crude oil which does not flow easily. It is referred to as
"heavy" because its density or specific gravity is higher than that of light crude oil. Characteristics of heavy oil:
 Gravity: very dense, between 7 and 25° API
 Viscosity (at reservoir conditions): highly viscous, from 10 up to over 10,000 centipoises

Table 1.1 Categories of heavy oil

Categories Properties
A Class Gravity < 25 API
Mobile Heavy Oil 10 < Viscosity < 100 cP
B Class Gravity < 20 API
Extra-Heavy Oil 100 < Viscosity < 10,000 cP
Non Mobile C Class Gravity < 12 API
Oil sands - bitumen Viscosity > 10,000 cP

One of the major mechanisms of oil recovery by thermal processes is due to the viscosity reduction of oil as
temperature increased. By reducing the oil viscosity a higher mobility is developed and the fluid flows much more
easily and/or is displaced through the reservoir. (The mobility of a fluid in the porous space of an oil reservoir is
7
expressed as the ration of the effective permeability to fluid viscosity.) Oil production is also inversely proportional
to viscosity. Raising the reservoir temperature has the effect of decreasing the crude oil's viscosity, sometimes
dramatically, and large increases in the production rate can be predicted. A useful correlation of oil viscosity as a
function of temperature and API gravity has been made by Farouq Ali and Meldau (1983) using viscosity-
temperature data of 60 heavy crude oil samples ( See Figure 1.3 ).

Figure 1.3. Oil viscosity as a function of temperature and gravity

According to this graph, the oil viscosity 𝜇 vs 1/T on semi-log coordinate is a straight line so the following
equation will be used.

𝐥𝐧 𝛍 = 𝐀 + 𝐁/𝐓 (1.1)
Where

T  absolute temperature
A and B are two constant to be determined by experimental data.

At the same temperature the viscosity of the crude increases with its density on a doubled logarithmic scale. As a
result, for the same increase in temperature the reduction of the crude oil viscosity is more evident for heavier
crudes. Data from some typical heavy oil reservoirs in the US and Canada are given in ( Figure 1.4 ).

8
Figure 1.4. Typical heavy oil Viscosity vs Temperature

For example, raising the reservoir fluid temperature to 300 °F (147 °C) will decrease the viscosity of the Cold Lake
"crude" (10 °API) from 100,000 cp to approximately 30 cp, or 10,000 times. The same increase in temperature
reduces the viscosity of the Kern River "A" oil reservoir (14 °API) only 500 times, from 2000 cp to 10 cp.

1.7 Problems in Applying Thermal Processes

The main technical problems associated with thermal techniques are poor sweep efficiencies, loss of heat energy
to unproductive zones underground, and poor injectivity of steam or air. Poor sweep efficiencies are due to the
density differences between the injected fluids and the reservoir crude oils. The lighter steam or air tends to rise
to the top of the formation and bypass large portions of crude oil. Data have been reported from field projects in
which coring operations have revealed significant differences in residual oil saturations in the top and bottom
parts of the swept formation. Research is being conducted on methods of reducing the tendency for the injected
fluids to override the reservoir oil. Techniques involving foams are being employed. Large heat losses continue to
be associated with thermal processes. The wet combustion process has lowered these losses for the higher-
temperature combustion techniques, but the losses are severe enough in many applications to prohibit the
combustion process. The losses are not as large with the steam processes because they involve smaller
temperatures. The development of a feasible downhole generator will significantly reduce the losses associated
with steam injection processes. The poor injectivity found in thermal processes is largely a result of the nature of
the reservoir crudes. Operators have applied fracture technology in connection with the injection of fluids in
thermal processes. This has helped in many reservoirs. Operational problems include the following: the formation
of emulsions, the corrosion of injection and production tubing and facilities, and the creation of adverse effects
on the environment. When emulsions are formed with heavy crude oil, they are very difficult to break. Operators
need to be prepared for this. In the high-temperature environments created in the combustion processes and
when water and stack gases mix in the production wells and facilities, corrosion becomes a serious problem.
Special well liners are often required. Stack gases also pose environmental concerns in both steam and
combustion applications. Stack gases are formed when steam is generated by either coal- or oil-fired generators
and, of course, during the combustion process as the crude is burned.

9
Chapter 2

IN-SITU COMBUSTION

2.1 Concept of ISC

In the porous rock of an oil reservoir, the oil can be ignited around the wellbore by means of an igniter or by a
spontaneous reaction of the oil to the air injected into the formation.
A burning front is built up, and the combustion is sustained by continuous injection of air or oxygen enriched air.
A small portion of the oil in place is burned furnishing heat to the rock and its fluids. The heat generated does the
below:
• Reduces the viscosity of the oil, increasing its mobility.
• Increases sweep efficiency and reduces oil saturation.
• Vaporizes some of the liquids in the formation generating steam and hot gases.
• Produces miscible fluids by condensation of the light components of the vaporized oil.

Figure 2.1. In situ combustion drive mechanism


The continuous injection of air or oxygen-enriched air develops efficient pressure maintenance and/or gas drive
mechanisms. As a result the oil is more easily displaced to the well producers by the slow movement of the burning
front through the reservoir rock.
The idea of in situ combustion was patented in 1923 by Wolcott and Howard. The first field test attempts to ignite
oil in a reservoir were conducted after 1930 in the Soviet Union and in the United States. Negative results were
reported because of the well's reduced injectivity. The field tests were started again in the United States during
1952 and extended after the first laboratory results were published in 1953 and 1954 by Kuhn and Grant.

10
2.2 Assets and Liabilities of In-Situ Combustion Process

Compared to other improved oil recovery processes, in-situ combustion is a highly complex process. This
complexity was not well understood by most early day in-situ combustion operators. This resulted in a high rate
of project failures in the 1960s, and contributed to the misconception that ISC is a problem prone process with
low probability of success. As a result, operators’ interest in the process waned, as reflected by the number of
project implementations since the 1970s. The truth, however, is that ISC is an attractive oil recovery process and
is capable of recovering a high percentage of the oil-in-place, provided the process is designed correctly and
implemented in the right type of reservoir.
Like other oil recovery processes, ISC has its assets and liabilities and no general criteria can be specified to assure
its success. The probability of an ISC project failure, however, can be minimized by recognizing its limitations and
designing the project accordingly. In this section we enunciate the advantages and limitations of ISC process and
examine critically the reasons cited for its declining popularity.

2.3 Assets of In-Situ Combustion Process

ISC is a unique oil recovery process. It can be viewed as a combination process. It encompasses some aspects of
nearly every known oil recovery method. These include steam distillation, steam displacement, CO2 flood,
hydrocarbon miscible flood, immiscible gas (N2) displacement, and water (hot and cold) flood.
Next to waterflooding, ISC is perhaps the most widely applicable improved oil recovery technique. The major
assets of ISC include the following:
 Thermally, it is the most efficient oil recovery process.
 It uses air, the least expensive and the most readily available fluid as injectant.
 ISC can recover oil economically from a variety of reservoir settings. The process has proven to be
economical in recovering heavy oil (10-20°API) from shallow reservoirs (less than 1,500 ft.), and light oil
(>30°API) from deep reservoirs (11,000 ft.).
 Though most combustion projects are implemented in heavy oil reservoirs, it is increasingly being used
to recover light oil from deep reservoirs. In the U.S.A, more combustion projects are in operation in light
oil reservoirs than in heavy oil reservoirs.
 It is an ideal process for producing oil from thin formation. Economically, successful projects have been
implemented in sandbodies ranging in thickness from 4—150 ft. The process, however, proved to be most
effective in 10-50 ft. sandbodies.
 Reservoir inhomogeneities have a less detrimental effect on the combustion process than on steam
injection process.
 Reservoir pressure has no effect on the technical success of the process. The process has been successfully
implemented in reservoirs ranging in pressure from vacuum to 4,500 psig.
 The formation permeability has minimal effect on the process. The process has been successfully
implemented in formations whose permeability ranges from 5 md to 10,000 md.
 Recovery efficiency is better than other oil recovery processes (see Table 2.1).
 The process can be implemented as a follow-up to waterflood and steamflood processes.
 The process can be applied in reservoirs where waterflood and/or steamflood is not effective. For
example, in deep reservoirs (greater than 10,000 ft.), steam injection is not effective due to excessive
heat losses and high lifting costs rendering waterflood economically unattractive. In such reservoirs,
combustion and gas are the only processes that can be applied to recover oil economically.
 Combustion projects permit the use of wider well spacing and can result in higher ultimate oil recovery
in comparison to steamflood.

11
Table 2.1. Recovery Efficiency of In-Situ Combustion Compared to Other EOR Methods

Process (A) (B) (C) (D)


Process Areal Sweep Vertical Compound
Displacement Efficiency Sweep Recovery
Efficiency % Efficiency Efficiency
% % %
In-Situ Combustion 95 70 85 56
Steamflood 65 70 85 39
Cyclic Steam — — — 20
Micro-Emulsion Flood 90 70 80 50
C02-Waterflood 80 50 80 32
NaOH-Waterflood 35 70 80 20

NOTE: D = (A) x (B) x (C)

Volumetric sweep efficiency = (B) x (C)

2.4 Laboratory Studies

In situ combustion mechanisms are largely a function of oil composition and rock mineralogy. The extent and
nature of the chemical reactions between crude oil and injected air, as well as the heat generated, depend on the
oil-matrix system. Laboratory studies, using crude and matrix from a prospective ISC project, should be performed
prior to the design of any field operation.

2.4.1 Oxidation Cells


The oxidation cells are used to obtain information regarding the reactivity of different oils in a porous medium
and the mechanism of the reactions.
The oxidation cell, into which air is injected continuously, is a core sample saturated with oil and heated to 940
°F. The temperature in the core sample, which is increased linearly with time, is recorded by thermocouples, the
oxygen is consumed, and the effluent gases are analyzed and measured.

In general, a reactive oil exhibits two successive oxidation reactions as shown in (Figure 2.2).

1. The low-temperature oxidation reaction (T < 600 °F) affects the lighter components of the crude.
Small amounts of CO2 and CO are formed even though 02 consumption is high. The oxygen remains
incorporated in the hydrocarbon chain which is not destroyed by combustion (Burger and Sourieau,
1985). In a reservoir, the oxygen consumed in the first reaction is reduced because the oil is heated
by the combustion gases that flow downstream.

2. The high-temperature oxidation reaction takes place in the narrow zone of the reservoir closest to
the combustion front and affects the heavier components of the crude. It is more important than
the low-temperature reaction because the oil is cracked, resulting in volatile and gaseous fractions
and a residue of coke like deposits on the sand grains.

The residue of coke like deposits, which constitute the principal fuel source, is burned to maintain the combustion
front.
The results obtained by the oxidation cell experiments also showed the reactivity of different oils. For example, a
12
paraffin-base and high-API-gravity oil could be completely

Figure 2.2. Differential thermal analysis of a crude oil (From Burger and Sahuquet,1972)

flushed out from the path of the combustion zone by the hot fluids, leaving no coke like deposit behind (Van
Poollen, 1980). The asphaltic and naphthenic base crude oils have aromatic compounds and therefore form the
greatest amount of coke. They are good candidates for the in situ combustion process because it is this coke like
residue which sustains the combustion front.
It also has been pointed out that the formation of coke is improved by the catalytic effect of different metals such
as nickel, vanadium, chromium, and so on, and by a larger specific area of the porous rock (matrix), as in a clay
sand. These conditions may give a light oil sufficient fuel availability to ensure the propagation of a combustion
front.

2.4.2 Combustion Tube Studies


Combustion tube experiments are mandatory to determine the parameters needed to design and implement field
projects. These data are used to make predictions of field test performance. As Sarathi1 points out, “Combustion
tube studies are the necessary first step in the design of an ISC project.”
Combustion tubes aim at representing a small volume of the reservoir. They are usually packed with native
reservoir cores or representative samples of matrix material and oil, placed in vertical position to minimize gravity
effects and heated to reservoir temperature. Ignition is usually started at the top by electrical heaters and the
combustion front is propagated downward. This allows propagation of a combustion front and the associated
chemical reactions at conditions close to those in a reservoir.
Temperature profiles, pressures, gas and liquids injection and production rates, and composition histories at the
inlet and outlet are recorded. ISC tube runs are unscaled and direct correlation of combustion tube results to the
field is not possible. However, as long as the runs are performed with reservoir rock and fluids at reservoir
conditions, the reactions of fuel deposition and combustion will be similar in both tube and reservoir. Tube runs
will not provide information on ISC sweep efficiency. They adequately model the chemistry of the process but not
the flow behavior in the reservoir, and only partially model the heat transfer processes. Flow behavior in the
reservoir is affected by gravity override, well spacing and geometry and reservoir heterogeneities, and tube runs
cannot reproduce these phenomena. Heat transfer from the tube to the surroundings can be much higher than
reservoir heat losses.
Two different schools of thinking exist on this heat transfer problem. Many experimenters use strip heaters
around the tube to lower the temperature gradient between the tube and the surroundings. This reduces heat
losses and allows front propagation at fluxes similar to those in the field. It can, however, lead to overestimation
of water/oil ratios in wet combustion if the strip heaters provide too much energy to the system, as they often

13
do. Information on front cooling by injected water may also be masked by the heaters. As a result, the extent of
the steam plateau may not be correct. Most of these types of experiments are bulky and time consuming and
require extensive instrumentation.
The other solution is to increase the air flux and minimize heat losses by insulation alone. This may slightly
overestimate air requirements and fuel content but is much simpler and easier to operate. As a result, it is widely
used.
The information that can be acquired from tube runs includes:

 Fuel burned
 Air required to burn a unit volume of reservoir
 Atomic H/C ratio of burned fuel
 Excess air and oxygen utilization
 Air/Fuel ratio
 Oil recovery from the swept zone
 Optimization of water/air ratio in wet combustion
 Composition of produced fluids
 Front temperature and stability

This last information is quite important in heavy oils to determine if the process is operating properly in the desired
high temperature regime. If high temperature cannot be achieved in ideal laboratory conditions it is likely that
field results would be worse.

2.5 Qualitative Description of In-Situ Combustion

The mechanism of oil displacement by a combustion front is complex. On a cross section made between an
injection well and a producer in a reservoir formation with uniform permeability, the combustion front has
practically an elongated S shape. The density differences between injected air and reservoir liquids make evident
the gravitational segregation and the tendency of gas to override (Figure 2.3).
This tendency is accentuated when the reservoir thickness is larger and is attenuated when the formation is
inclined. Under steady-state conditions the various zones formed in the reservoir during the in situ combustion
process are as follows:

Zone 1 Is the combustion front, where the oxygen is consumed in burning the coke
deposited on rock grain surfaces and steam is one of the products formed. Zone 1
has the highest temperature, from 600 to 1200 °F

Zone 2 Is left behind by the combustion front as a hot and clean sand that heats the injected
air before the air reaches the front.

Zone 3 Is the vaporization zone ahead of the combustion front, where the lighter
hydrocarbons and the interstitial water are vaporized and the heavier hydrocarbons
are thermally cracked, leaving the coke like deposit on the sand grains.

Zone 4 Is the condensation zone where the steam and the hydrocarbon gases move
forward into the cooler reservoir, condense, and a large amount of heat is released.
The oil displacement is increased by the oil's lower viscosity, higher mobility, and
the miscible effect of the mixture between the condensed gas and the oil bank.

14
Zone 5 Is characterized by a water saturation higher than the interstitial water saturation
(water bank) which pushes the oil bank ahead to the producers (Latil, 1980).

Figure 2.3. Schematic representation of in situ combustion process and the various zones as formed in the oil
reservoir

The knowledge of the displacement mechanism of the in situ combustion process makes it possible to improve
its efficiency. For instance, considering the tendency of gas to override, the injectors are perforated only in the
lower half of the pay zone (limited entry) and the producers avoid the top of the formation (Figure 2.3).
Since the specific heat of the injected air is too low to carry the heat accumulated in the burned zone, more than
half of the heat generated remains behind the combustion front
(zone 2). The heat is lost by conduction in the adjacent formations. To achieve a better thermal efficiency, after
starting off the process as a dry combustion, water is injected in combination with the injection of air. This process
is called wet combustion.

2.6 Mechanism under Processes of In-Situ Combustion Working

This section includes the processes of in-situ combustion used for the displacement of heavy oil from reservoirs.
Basically, there are three different processes of in-situ combustion: forward combustion, wet combustion and
reverse combustion.

15
Figure 2.4. Schematic illustration of the in-situ combustion process. (After Nelson and McNeil 1961; courtesy of
the Oil and Gas Journal.)

2.6.1 Forward Dry combustion


The forward combustion process is schematically shown in Fig. 2.4. In this process, the reservoir is ignited in the
vicinity of an air injection well. Moreover, the combustion front propagates away from the injection well. The
continuous injection of air maintains and drives the combustion front through the reservoir; in general, towards
the production wells or the direction of air flow. In addition, if the air injection is reversed (e.g.. air injection at the
injection well is stopped and then started at the producing well), the direction of flow of the oil bank will be
reversed (e.g., in the direction of air injection). The oil bank is thus forced to move through the burned zone,
whereas the flow pattern of combustion front will remain unchanged (Fig. 2.5).
The temperature distribution in a reservoir during the forward combustion process is shown in Fig. 2.5. It is evident
that the temperature and heat content of the burned zone are quite higher than those of the rest of the reservoir.
The oil, therefore, will be heated to temperatures ranging from 500° to 700°F, which results in the reduction of
oil viscosity by several orders of magnitude. This reduction in oil viscosity, therefore, allows high-viscosity crude
oils to flow freely through the reservoir rock towards the production well. The forward combustion is also known
as dry combustion. Moreover, several variations of the forward in-situ combustion have been suggested to
enhance the efficiency of this process. In several laboratory and field test, mainly two variations, i.e., wet and
partially quenched combustion were employed. The pilot test operated by the Cities Service Company
demonstrated that dry combustion could recover the low-gravity crude oil economically, whereas laboratory
studies showed that wet combustion would be superior to dry combustion.

Figure 2.5. Temperature distribution in the reservoir during a fireflood. (After Nelson and McNeil, 1961; courtesy
of the Oil and Gas Journal.)

2.6.2 Forward Wet Combustion


The wet combustion process, also known as COFCAW (combination of forward combustion and waterflood)
(Parish and Craig, 1969), transfers the accumulated thermal energy forward (upstream) from the combustion
front. Water is the transferring agent because of its high heat capacity and its latent heat of vaporization.

16
Figure 2.6. Temperature profiles in dry and wet combustion

Air and water are injected concurrently or alternately into the injection well. The injected water flashes into
superheated steam, passes through the combustion front, and transfers heat to the area ahead of the front.
A comparison of the temperature profiles shows that the wet combustion process operates at lower temperatures
behind and at the combustion front. Ahead of the combustion front the hot zone (the steam plateau) is increased
in size (Figure 2-6).
The advantages of the process are evident: A much larger area of oil saturated rock is affected by higher
temperatures, oil mobility and sweep efficiency is increased, less fuel at the combustion front is necessary, and
less air is required to sweep the reservoir. in oil reservoirs with low permeability and higher content of dirty sand
(swelling clay), the introduction of water into formation may reduce injectivity and increase air injection pressure.

2.6.3 Comparison of wet and dry combustion


Figs. 2.5 and 2.6 show that much heat is left in the burned zone behind the combustion front. During the dry
combustion process, this heat is dissipated through the base rock and reservoir cap. Moreover, when water in a
moderate amount is injected simultaneously with the air, it converts into superheated steam in the reservoir. The
steam travels a short distance from the injection well, whereas the water evaporation front remains close to the
injection well (Fig. 2.7a). As the mixture of superheated steam and air approaches the combustion front, only
oxygen is utilized in the burning process. Upon passing the combustion front, the superheated steam mixes with
nitrogen from air, and the flue gas phase consists mainly of the mixture of carbon monoxide and carbon dioxide.

17
Figure 2.7. Comparison of wet and dry combustion

The flue gas phase displaces the oil in front of the combustion zone and condenses as soon as its temperature
drops to nearly 400 F. The size of the steam zone depends on the amount of heat recovered from the burned zone
upstream. In order to bring the evaporation front as close as possible to the combustion zone, water in sufficient
amount is required for injection (Fig. 2.7b), which, in turn, would result in transporting major part of excess heat
of the burned zone to the steam zone. The size of the steam zone, therefore, would increase, and for a given well
spacing the combustion front would travel a much shorter distance than that for dry combustion. This would
result in displacing more oil towards the producing site. It has been shown in the laboratory that optional wet
combustion requires about one-third of the air needed for dry combustion.
In order to compare the dry and wet combustion processes, a field test was performed by the Cities Service
Company in 1974 (Joseph and Pusch, 1980). It was observed that the air volume required to process a specific

18
volume of reservoir reduced to 63% with wet combustion as compared to dry combustion. As the major portion
of the costs for a combustion project is attributable to maintenance and operation of the air compressor, wet
combustion offers significant economic Incentive. Moreover, the time required to process pattern volume will
also be decreased due to less air needed per pattern in wet combustion. A realistic comparison of the two recovery
processes on the basis oil production, however, was very difficult due to measurements and allocation problems.
In addition, a direct method of comparison can be developed from the study presented by Gates and Ramey
(1980). Their technique showed that the oil production for a given field operation can be predicted from the
relationship between the volume of oil recovered and that burned. If such a relationship is valid, then two
combustion processes can be estimated by comparing the burned volumes generated by each process.
The selection between dry combustion and wet combustion is an important decision in conducting a field project.
Several laboratory results indicated that the injection of water either simultaneously or alternatively with air in
the ignition wells reduces the air/oil ratio. Based on data of various field projects, the correlation between air/oil
ratio and water/air ratio was observed to be statistically insignificant in the presence of widely varying properties
of the reservoirs (Chu, 1982). The field test of wet and dry combustion conducted by Cities Service Company
(Joseph and Pusch, 1980) yielded the following conclusions:

1. The volumetric sweep was improved to a great extent with wet combustion as compared to dry combustion
which, in turn, indirectly implies an increase in oil recovery.
2. The project life is significantly reduced, which means improving the economics.
3. The reduced air requirement in wet combustion for a specific volume of reservoir decreases the operating
cost and improves the economic future.

It should be noted that these encouraging results may not be entirely applicable to other reservoirs having
different characteristics and conditions. The possible advantages of wet combustion over dry combustion,
however, can he examined by numerical simulation, laboratory experiments and pilot field tests.
The partially quenched combustion is schematically shown in Fig. 2.7c. When water is introduced in insufficient
amounts into the combustion zone, the partially quenched combustion would occur. In a laboratory test, it has
been shown that oxygen is utilized within 0.4 - 3.3 ft at a temperature of 400° F. Thus, the introduction of water
in the combustion zone would partially quench the combustion, allowing the oxygen to travel until it comes in
contact with oil at 400°F. The combustion zone, therefore, travels at the speed of the cooling water. In this way,
all the processes such as partial evaporation and heating of water, heat production by combustion, and heat
recovery from the formation occur in one front, which travels fast (Fig. 2.7c).

2.6.4 Reverse Combustion


When oil is too viscous to flow under reservoir conditions but the reservoir has an adequate air permeability, it is
possible to produce oil by reverse combustion. In this case, the combustion front moves counter to the air flow
(Figure 2-8). After ignition, the well is put into production and another well is used for injection. The front moves
in the same way in which a cigar is consumed—by expelling the air instead of inhaling it (Crawford, 1971). The
process has limited use and was tested in a tar sand reservoir in Bellamy Field (Trantham and Marx, 1965).

19
Figure 2.8. Temperature profile in reverse combustion

The reverse combustion process is schematically shown in Fig. 2.9 In this process, air is injected through ignition
wells that eventually become oil-producing wells. Initially, the reverse combustion process starts as a forward
combustion process. After the burning zone moves to within a short distance from the ignition well, air injection
is stopped in the ignition wells, and is started in adjacent wells. The air injection is continued in the adjacent wells
in order to drive the oil towards the wells which previously were ignition wells. The combustion front moves in
the opposite direction towards the adjacent wells (Fig. 2.9). The oxygen required for combustion is only supplied
by the air, which is continuously injected in the adjacent wells. In addition, if the oil in the adjacent wells ignites
spontaneously, the air injection for reverse combustion is stopped, and the process essentially converts to the
forward combustion process.

Figure 2.9. Schematic illustration of the reverse combustion process. (After Berry and Parrish. 1960; courtesy
of the Society of Petroleum Engineers of AIME.)

20
Figure 2.10. Variation of ignition time as a function of temperature for several air-injection pressures. (After
Dietz and Weijdema, 1968; courtesy of the Producers Monthly.)

Fig. 2.10 It has been shown that the ignition occurs spontaneously within a period of a few days, except for low-
reactivity crudes, i.e., oils at near arctic reservoir temperatures and low reservoir pressures (Dietz and Weijdema,
1968). Fig. 2.10 illustrates the results of ignition time as a function of reservoir temperature for crude oils having
different reactivity. These results probably explain the reason why several pilot field tests of reverse combustion
process have failed, even though the laboratory experiments have shown satisfactory results.

2.7 Operating Practices

In addition to the standard field equipment for oil production, ISC requires particular attention to air compression,
ignition, well design, completion, and production practices.

2.7.1. Compressors
Air compression systems are critical to the success of any ISC field project. Failures of the past can often be traced
to poor compressor design, faulty maintenance or operating mistakes.
The factors to be considered when selecting compressors include peak air requirements, injection pressure,
capital cost, power requirements, operation and maintenance costs and other relevant technical and economic
parameters specific to the field considered. Compressor terminology varies among manufacturers. It is best to
obtain a complete description including compressor, driver, interstage cooling system, and all ancillary equipment
including control and safety systems from each vendor being consulted.
Air compression causes high temperatures because of the high cp /cv ratio of air.
Compressor design must consider these high temperatures to ensure continuous, sustained operations free from
the corrosive effects of air and the explosion hazards of some lubricating fluids. Mineral oils are not
recommended. Synthetic lubricants withstand the higher temperatures and offer lower volatility and flammability
21
than conventional lubricants.

2.7.2 Air Requirement


The fuel depositional characteristics of the reservoir oil are the most basic parameters in designing a fire-flood.
Coke deposited as fuel is measured in pound per cubic feet of reservoir rock. If this value is too low, combustion
cannot be supported. If it is too high, flame movement is too slow because all fuel must be burned before the
flame advances. Fuel deposition also determines the amount of air required to advance the flame through a set
volume of reservoir rock. The amount of air required to bum through one cu ft of oil sand as a function of oil API
gravity, which shows how fuel deposition varies with API gravity of reservoir oil.
The main factors which govern the volume of air required for in situ combustion are: amount of fuel supply (coke)
of the oil being burned and the efficiency of oxygen utilization.
As the combustion zone moves through the reservoir, it continuously emits heat. The heat moves in the forward
direction by conduction, as sensible heat in liquid and gas and as heat of vaporization in vapor. The hot fluids flush
out volatile and mobile substances from the path of the combustion zone leaving behind residual hydrocarbon
(coke), which is the fuel supply for combustion. The combustion zone can only move as fast as it depletes the
coke. If the amount of coke deposited by the crude oil is excessive, the rate of advance of the combustion zone
would be slow and the air requirement would be large. On the other hand, if the oil is paraffin-base and of high
API gravity it could be completely flushed out

2.7.3 Ignition
Ignition and maintenance of high combustion temperatures, especially in heavy oil projects, are the most critical
factors of an ISC project.
Ignition can occur spontaneously if the oil is reactive, the reservoir temperature high enough, and the reservoir is
reasonably thick. Various models have been proposed to determine the time for spontaneous ignition.
When spontaneous ignition does not occur or is not desired (i.e., in heavy oil reservoirs where it is important to
maintain high combustion temperatures), the most appropriate ignition method to use depends on the reservoir
and the equipment available on site.
Down hole gas-fired burners allow good control of the temperature of injected gases and may be operated at a
greater depth than other methods. The disadvantages include the need to run multiple tubing strings in the
injection wells. Some particulates, such as soot, may be carried into the formation if the gas does not burn cleanly.
Catalytic heaters run at lower temperatures but are sometimes prohibitively expensive. Electrical heaters can be
lowered with a single cable, and can provide excellent temperature control. They can be reused repeatedly. There
is, however, a depth limitation because of electrical power losses in the cable.
Chemically enhanced ignition does not have a depth limitation but may require handling and storage of dangerous
materials. Fuel packs are not recommended because of poor temperature control and no uniform ignition across
the entire reservoir thickness. Well damage from elevated temperatures and plugging by particulate matter may
occur.
Steam may be used to locally increase reservoir temperature and facilitate auto ignition. It suffers from depth
limitation because of wellbore heat losses, but when the conditions are right it can be a very simple and effective
method for ignition.

2.7.4 Well Design and Completions


ISC wells must be designed to account for several factors amplified by the combustion process, namely high
temperature, corrosive environment and sand and clay control. Safe operations should be the primary concern.
Typical well designs for injection and production are shown in Fig 2.11. Completion type and design depends on
the reservoir being considered. Laboratory testing for sand control and completions can help to determine the
best completion technique for a given field. Care has to be taken to properly cement the wells. There are cement
formulations that are stable at high temperatures. Open hole completions may be used in conjunction with slotted
liners, screens, gravel packs or various other sand and clay control methods. To maximize productivity, producing

22
wells should be completed toward the bottom of the zone of interest to take advantage of gravity drainage and
avoid hot gases as long as possible. Rat holes have been used successfully in certain heavy oil combustion projects
to increase the effect of gravity drainage.

2.7.5 Injection and Production Practices


Safe air injection requires that the surface injection equipment and the injection well are free of hydrocarbons.
All lubricants used in compression and down hole operations should be synthetic or non-hydrocarbon types. All
equipment, tools, lines, tubing, work strings and injection strings must be clean and hydrocarbon free. Personnel
at all levels should be aware of the importance of preventing hydrocarbons in the injection wells. As a safety
measure to protect injection wells if the compressor is shut down, a system to prevent backflow of oil from the
formation must be present at every injection well.
Down hole temperatures in producing wells increase as displaced oil, hot water and steam fronts reach the well.
Producers are preserved by down hole cooling and proper material selection. Figure 2.12 provides an estimate of
the water requirements to maintain bottom hole temperature no higher than 250ºF as a function of oil and water
production rate and formation flowing temperature. Significant additional oil recovery can be obtained from hot
wells with down hole cooling, especially if the well is completed in the lower section of the producing zone to
maximize gravity segregation in the reservoir. In many cases, after the combustion front has moved through the
well it is possible to convert the former producer to a new air injector, thus realizing significant cost reductions
over the life of the project.
Monitoring is crucial for proper combustion operations. In addition to testing individual producers for oil and
water rates, injected fluids must be measured. Also, produced gases must be measured and analyzed to
determine the efficiency of the combustion operation. Down hole temperature measurements are essential to
calculate the size and location of the burned zone. Flow line temperatures can indicate thermal stimulation or
down hole problems.
Combustion projects generate waste water, flue gases and pollutants from compression and oil handling
equipment. Local pollution disposal regulations must be consulted prior to the design of any in situ combustion
operation.

23
Figure 2.11. Typical production well design

Figure 2.12. Water needed to cool hot wells

24
Table 2.2. Application of Pollution Control Systems to a Fireflood Project

Equipment Gas Treated Pollutant Method Suggested


Removed Application

Flare stack Flue gas None Vent to Flue gas meets air
atmosphere quality regulations

Flare stack with flame Flue gas H/C(1) ,CO(1) , S-gases Burn Flue gas with enough
(1,2)
burner H/C to support
combustion
(>200Btu/Scf)
Thermal Flue gas H/C, CO, S-gases Burn Flue gas not suitable for
(1,2)
incinerator a flare with burner or a
catalytic incinerator
(heat value > 85 Btu/Scf
but < 200 Btu/Scf)

Catalytic Flue gas H/C, Co, S-gases Burn Flue gas with heat
(2)
value < 85 Btu/Scf
incinerator (excess air)

H 2 S scrubber Flue gas (4) H2 S Chemical reaction Flue gas containing H 2 S


but

acceptable amounts of
other S gases
SO 2 scrubber Incinerator (4) SO2 Chemical reaction When H2 S removed
exhaust from flue gas is
inadequate or
impractical
1. Removal efficiency may be poor

2. S-gases are converted to SO 2

3. May increase amount of CO

For flue gas with heat values > Btu/Scf

4. Typical removal efficiency is 90-95%.

In general, environmental problems are similar to those posed by steam injection. The produced water may
contain H2S and/or CO2 which may require special handling and anti-corrosion equipment. Flue gases may
contain hydrocarbons, H2S, CO2, CO and other trace amounts of sulfur gases. Table 2.2 summarizes the various
pollution control systems suitable for combustion projects and their recommended applications. Sarathi1 also
provides detailed descriptions of the various types of systems and their uses. Some other problems that can be
encountered are sand production, corrosion, emulsions, well failures or compressor failures.

25
2.8 Pilot Test

Prior to field development of in situ combustion, pilot tests involving a small portion of the reservoir are conducted
in order to check the laboratory results and to obtain

- Data to design surface injection facilities (injection pressure, injection rate).


- Information regarding formation damage, sand invasion, emulsions, corrosions, and so on.
- Data about the existence of preferential fluid flow, directional permeability, gas fingering, impermeable
barriers, and so on.
- Production rates, temperatures, and effluent gas measurements; burning front configuration and velocity,
and so on.
- Information regarding the oil recovery (only if the field pilot is considered a confined pattern).

Figure 2.13. (a) Open inverted and (b) confined direct five-spot

If the combustion is performed in four adjacent inverted five-spot patterns with injection wells Il , I2 , 13 , and I4
(Figure 2-13b), the four injection wells then encompass a direct five-spot pattern which may be considered a
confined pattern with a central production well P.
The cumulative oil production of the well P divided to the initial oil in place, corresponding to the direct five-spot
pattern area, gives an estimate of the oil recovery.

2.9 Patterns Sweep, Invasion and Displacement Efficiencies

Two of the most important facts to know in order to control and evaluate an in situ combustion process are the
volume of the burned zone and also the reservoir volume affected by heat and its carrying agents.

The specific notations defined by SPE (1986) are:

EA = areal efficiency (used in describing results of model studies only), is the area swept in a model divided by
total model reservoir area (see 𝐸𝑃 ).
𝐸𝑃 = pattern sweep efficiency (developed from areal efficiency by proper weighting for variations in net pay
thickness, porosity, and hydrocarbon saturation), is the hydrocarbon pore space enclosed behind the injected
fluid or heat front divided by the total hydrocarbon pore space of the reservoir or project.
𝐸𝐼 = invasion (vertical) efficiency, is the hydrocarbon pore space invaded (affected, contacted) by the injection
fluid or heat front divided by the hydrocarbon pore space enclosed in all layers behind the injected fluid or heat
front.

26
ED = displacement efficiency, is the volume of hydrocarbons displaced from individual pores or small groups of
pores divided by the volume of hydrocarbon in the same pores just prior to displacement.
𝐸𝑉 = volumetric efficiency, is the product of pattern sweep and vertical efficiencies.

EV = EP 𝑬𝑰 ( 2.1 )

EDb = displacement efficiency from the burned portion of the in situ combustion pattern (considered 100
percent).
EDu = displacement efficiency from the unburned portion of in situ combustion pattern.
ER = overall reservoir recovery efficiency, is the volume of oil recovered
Divided by the volume of hydrocarbon in place at the start of the project.

ER = 𝑬𝑷 𝑬𝑰 ED = 𝑬𝑽 ED ( 2.2 )

The scheme of a combustion front advancing in an oil reservoir is given in Figure 2-14 with an explicit presentation
of the notions just defined (Burger and Carcoana, 1975).
The burned and unburned zones, the pattern sweep efficiency and the vertical efficiency appear clearly enough.
It is important to note that pilot tests in the field and in commercial scale development of in situ combustion
pointed to the existence of two unburned zones :

- Unburned zone I: inside the horizontal projection of the area covered by the combustion front; this zone
was not observed in the lab experiments.
- Unburned zone II: strictly speaking, the zone in front of the combustion zone.

Table 2.3. Pattern Sweep Efficiency

Field Pattern Sweep Field


Efficiency, %

Delaware Childers, Oklahoma ~100


South Belridge, California ~100
S. Oklahoma, Oklahoma 85
S.E. Kansas (Humbolt Chanute) 70
Shannon, Wyoming 43
Delhi, Louisiana 50
Suplacu de Barcau, Romania 85
A13 Sloss, Nebraska 50
Niitsu, Japan ~100

27
Figure 2.14. Unburned zones, pattern sweep, and invasion efficiencies

2.9.1 Sweep Efficiencies


The pattern sweep efficiency EP is determined approximately in the field by controlling the location of the
combustion front at a given time. For an enclosed five-spot well pattern the flow of injected air and combustion
gases relative to the movement of the oil corresponds to an infinite mobility ratio for choosing the areal sweep
efficiency. Different modeling techniques show in this case an areal sweep efficiency of 62.6 percent, limited by
practical considerations to 55 percent (Nelson and McNiel, 1961).
Some of the sweep efficiency values obtained in different five-spot pattern projects are presented in Table 2-4.
The higher values occurred because the inverted five-spot patterns were not confined.
The vertical sweep efficiency EI , is determined by taking post combustion core samples. These indicate thickness
of the burned zone and the amount of coke formed in the zone swept by the combustion front. It can also be
determined taking into consideration the fact that the resistivity of the burned zone is 2 to 10 times higher than
before combustion. The vertical sweep efficiency EI is largely dependent on the thickness of the formation and
has higher values for thinner pay zones.
The Delhi (Louisiana) project, where the thickness of the pay zone was only 2.5 ft, was characterized by a vertical
sweep efficiency EI value close to unity. In the North Tisdale project, core samples taken from a well drilled in the
area swept by combustion showed vertical efficiency EI = 27%. This time the burned portion of the formation
thickness was located in the middle of the pay zone and not in the upper half as usual. Other EI, values reported
ranged between 26 percent (South Oklahoma) to 50 percent (Trix Liz, Texas) and 60 percent (Emma Fry III, Illinois,
and South Belridge, California), with an average value of 35 to 40 percent obtained at Suplacu de Barcau
(Romania).

It is important to note that the burned zone is a clean sand with improved porosity and permeability.
The displacement efficiency EDu for the unburned zone l is known by the oil saturation values taken from core

28
samples. For instance, the North Tisdale project had a displacement efficiency EDu = 50% of the initial oil content,
and the South Belridge (California) oil saturation value within the unburned zone I was less than half of the initial
oil content. In the Suplacu de Barcau project the bottom coring of three wells drilled in the area swept by the
combustion front shows an average value of EDu = 43% (Turta, 1972, 1974).
The displacement efficiency EDu for the unburned zone II is very difficult to estimate. The oil from the unburned
zone II is influenced by heat conduction. This increases its mobility and thereby makes it more easily displaced by
the uncontrolled movement of the combustion front fluids. A bottom coring well drilled on a water-based mud in
the unburned area II in the North Tisdale combustion project showed the following oil saturations (Martin and
Alexander, 1971) :

40% at the top of the production formation


32% in the median zone
50% at the bottom of the pay zone

compared with 𝑆𝑜 = 60% at the start of the project. Assuming an average value of 40 percent, the displacement
efficiency in zone II is EDu = 33%.

2.10 Oil Consumed In-Situ

As a result of distillation and thermal cracking, the oil consumed in situ deposits on the rock the coke available for
combustion. The amount of oil (fuel) consumed is a function of the reservoir characteristics and of the volumetric
sweep efficiency, EV of the process. For instance, assuming reservoir rock porosity is 26 percent and oil saturation
at the start of the project is 70 percent, the oil consumed for a fuel content Cu = from 0.8 to 2.6 lbm/ft3 varies
between 5.8 and 19 percent for EV = 100% and between 1.7 and 5.7 percent for EV = 30%.

2.11 Oil Recovery by ISC

The value of the oil recovery reported from most of the in situ combustion experiments and projects ranges
between 40 and 60 percent of the oil in place at the start of the process. Higher recovery values are suspected to
be the result of oil coming from outside the pattern or zone taken into consideration.
The pattern and vertical sweep efficiencies EP and EI , are estimated using the procedures just described. The
displacement efficiency EDu for the unhurried zone I inside of the area swept by the combustion front is also
determined using these procedures. The displacement efficiency in the zone II outside of that area is unknown
and may be expected to be less than in the underlying zone. A range of possible oil recovery can be obtained by
considering
1. The minimum assumption, when the displacement efficiency is applied only to the unburned zone I (no oil
produced from the unburned zone II).
2. The maximum assumption, when the displacement efficiency is applied to both zones I and II.

For the unit volume of rock, the oil produced by in situ combustion in reservoir conditions (Burger and Carcoana,
1975) is

𝑵𝑷𝟏 = 𝑺𝑶 ∅ EDu EP (𝟏 − 𝑬𝑰 ) + (𝑺𝑶 − 𝑺𝑶 𝒄𝒐𝒏𝒔 ) ∅ EV (2.3)


𝑵𝑷𝟐 = 𝑺𝑶 ∅ EDu (𝟏 − 𝑬𝑽 ) + (𝑺𝑶 − 𝑺𝑶 𝒄𝒐𝒏𝒔 ) ∅ EV (2.4)

29
And the corresponding recovery factor is

𝑵𝑷𝟏 𝑵𝑷𝟐
𝑬𝑹𝟏 = 𝑺𝑶 ∅
and 𝑬𝑹𝟐 = 𝑺𝑶 ∅
(2.5)

The recovery factor ER is higher than 𝐸𝑅1 and less than 𝐸𝑅2 and refers to the oil in place at the start of the project.

2.12 Factors Affecting In-Situ Combustion

The success of any improved oil recovery program depends chiefly on reservoir considerations and the
combustion process is no exception. This despite is being the case; many early fireflood operators paid scant
attention to geology during the site selection process. Reservoir geological characteristics played a major role in
the outcome of many past combustion projects. Examination of the reservoir characteristics of the California,
Oklahoma and Texas fireflood projects (states that accounts for more than 70% of the implemented U.S. combus-
tion projects) indicate that the structure, lateral continuity, and physical characteristics of the individual sand
layers within the reservoir as well as the reservoir heterogeneities played a significant role in the performance of
these projects.
Lack of good sand continuity (due to complex lateral faces variations) and channeling have been cited as one of
the cause of failure of many California firefloods (Simm, 1967). Since in situ combustion is an interwell drive
process good horizontal continuity is critical to the success of the project. Gaps in formation overburden or leaky
interzonal seals in stratified reservoirs can allow fluid to leak into overlying strata and reduce the effectiveness of
the injecting. Fractures and joint trends, however subtle, may create preferential flow channels that influence
recovery efficiency.
Therefore, knowledge of the geologic characteristic of the site is important for the proper evaluation of a prospect
for in situ combustion. Earlougher, et al (1970) in analyzing the performance of the Fry in situ combustion project,
Illinois indicated that the reservoir geology played a prominent role in the outcome this project and stressed that
an understanding of the reservoir geology is essential to the design and successful operation of a combustion
project.
The objective of a geological reservoir description should be to provide a clear, concise picture of the qualitative
and quantitative parameters of the reservoir so that the engineer can design a scheme that most appropriately
matches the reservoir conditions.
The key geological parameters to be considered when selecting a site for a fireflood project include: the degree
and extent of lateral and vertical reservoir continuity, depth, thickness, structural attitude and dip, overburden
competence, reservoir heterogeneities, and of presence of gas cap and aquifer.
The discussion to follow outlines briefly the extent to which these and other rock-fluid parameters will impact a
fire-flood project.

2.12.1 Lateral and Vertical Extent of Reservoirs


The continuity of individual sand layers within the producing formation, especially in thin, lenticular sands is a
factor of major importance to the successful operation of fireflood. In-situ combustion require significantly more
capital investment per unit of production than water flood because of the need for ancillary equipment, such as
the air compressors, has high operating costs and is manpower intensive. This higher cost means that the volume
of oil in place per unit area must be above a certain minimum to make the project economically viable. In thin
reservoir, the total oil in place is a function of porosity, oil saturation and areal extent of the reservoir.
The success of the combustion projects implemented in the thin and often poor quality south Texas strand plain
barrier island (lagoon and near shore environment deposits) reservoirs [Glen Hummel and Gloriana (Buchwald, et

30
al, 1973), the North Government Wells (Casey, 1971), the Charco Redondo (Howard, et al, 1976), and the West
Casa-Blanca (Eskew, 1972)] is in part can be attributed to the excellent lateral continuity of the sands. In these
reservoirs the multiple, thin, blanket type oil column are more widespread and separated by shale stringers and
tight cemented mudstones. This made it an ideal geometry for achieving favorable sweep during combustion.
The failure of many 1960s California firefloods such as those undertaken in such fields as Ojai, White Wolf,
Placerita Canyon, Pleito Creek, and Tepusquet Canyon can also be attributed in part to the lack of reservoir
continuity. The formation at these sites, though, exhibit excellent porosity, good permeability and good oil
saturation composed of a series of overlapping sand lenses, separated by interbedded impermeable shale layers.
The poor lateral continuity and compartmentalization resulting from complex lateral faces variation did not permit
the free movement of fluids. The general lack of areal continuity and poor understanding of the geology of the
site by early operators contributed to the failure of the process in these reservoirs. Many of these pilots were
unfortunate in site choice because they were implemented in properties that did not prove to be economic as
primary producing cases.
Hence lateral and vertical extent of a reservoir is one of the key parameter to consider in the site selection process
for a fireflood project. The degree and extent of lateral and vertical reservoir continuity significantly affect the
performance of the combustion process. Clean, well-sorted sands tend to have good horizontal and vertical
continuity. Reservoir continuity can be reduced by disseminated finer grains, by the local occurrence of various
types of shale interbeds, and cementation materials. The study of setting in which the sands were deposited can
give an approximation of the reservoir continuity in lateral and vertical perspective. A complete characterization
of faces distribution would help to predict how flow barriers could affect reservoir performance.

2.12.2 Vertical Depth


Reservoir depth is not one of the restrictive parameters for the combustion process. Economically successful
combustion projects have been implemented in reservoirs ranging in depth from 300–11,500 ft. Depth, however,
is a factor in terms of temperature, pressure and well cost. Shallower depth (less than 200 ft.) would severely limit
the pressure at which air could be injected. With increasing depth, air injection pressure generally increases with
a corresponding increase in compression cost (larger compressor). Deeper reservoirs are usually hot enough, that
spontaneous ignition of in situ hydrocarbon is likely upon air injection.
Deeper reservoirs generally contain lighter oils and air injection at high pressure into these reservoirs can offer
some unique technical opportunities for improved oil recovery. Apart from combustion and the attendant oil
recovery by displacement, other reservoir mechanisms also contribute to oil recovery. These include: reservoir
pressurization, stripping of the light ends of in situ crude by the combustion gases, oil swelling and high-pressure
miscibility effects (Yannimaras, et al, 1991). With greater injection pressures greater injective capacity can be
obtained which in turn allows well spacing to be enlarged.
Well drilling and completion costs, however, increases with depth. Larger compressors are needed to meet the
injection pressure requirements. Larger compressors are more expensive to purchase, operate and maintain.
Depth also effects the fluid lifting costs, especially in wet combustion process. Thus economic considerations will
impose a practical upper depth limit. This may be on the order of 12,000–12,500 feet.

2.12.3 Reservoir Thickness


Sand thickness is one of the more important parameters for the combustion process. The large difference in
density between air and the reservoir fluids gives the air a tendency to override the oil column and consequently
bypass much of the oil if the reservoir exceeds a critical thickness. A thin oil sand tends to counter this override
tendency and favor a more uniform displacement and vertical sweep. In a thin heavy oil reservoir, rapid transfer
of heat to the bottom of the sand will permit combustion front to advance at the bottom more rapidly than it
would be possible in thick sand (Boberg, 1988). If the sand, however, is too thin high overburden heat losses may
drop the temperature below that necessary to sustain a combustion front and can lead to low temperature
oxidation and loss of recovery. Preferably pay thickness should at least be four feet and should not exceed 50 feet.

31
It is preferable that very thin reservoirs (less than 8 ft. thick pay) considered for fireflood be contain multiply
stacked thin sand separated by non communicating vertical barriers to take advantage of heat conduction to the
vertical direction. This can not only minimize the heat losses to the overburden but can also aid in promoting and
sustaining high temperature combustion mode in heavy oil reservoirs. The Fry fireflood in Illinois (Hewitt and
Morgan, 1965) and the Gloriana in situ combustion project in Texas are examples of successful projects
implemented in multiple thin (less than 5 ft. thick) sand reservoir.
Formation thickness is also an important consideration in reservoirs containing oil not readily susceptible to auto-
ignition. In such reservoirs the near well-bore area must be heated to a high temperature to initiate ignition. If
the formation is very thick (> 50 ft.) the amount of heat needed to raise the well-bore vicinity above the oil’s auto-
ignition temperature can be very large and expensive. Formations up to 60 feet thick have been ignited using
artificial ignition techniques.

2.12.4 Structural Attitude and Dip


Structural attitude and dip are important consideration in the location of wells for a combustion project. Injected
air and combustion front movement will be more rapid toward up dip wells than toward wells low on the
structure. In dipping reservoirs it is advisable to locate the air injector’s down-dip and production wells up the
structure to compensate for the expected flow of air up dip. In steeply dipping reservoirs some operators
preferred injecting air at the top of the structure to take advantage of gravity in the recovery of hot mobile crude
affected by combustion (Gates and Skalar, 1971). In the steeply dipping Webster reservoir, the combustion project
was initiated as a crestal drive, in part, to heat the oil at the top of the structure and promote migration of oil
toward the flanks of the anticline (Soustek, 1994). Dip and the resulting gravity dominance played a major role in
the economic success of the Santa Fe Energy CO’s (now part of Texaco) Midway Sunset Combustion project.
Turta (1995) recommend locating the ISC pilot at the uppermost part of the structure. The reason behind this
recommendation is that the burned volume of the pilot located at the upper part of the reservoir can be more
accurately be determined and both the air-oil ratio (AOR) and the incremental oil recovery due to combustion can
be estimated more reliably. Also by locating the pilot up dip, restoration of the burned zone can be avoided in
case the air injection is terminated due to compressor failure.

2.12.5 Overburden Competence


The producing formation at the project site must have sufficient and competent overburden so as to confine the
injected air within the pay zone. Gaps in oil sand overburden or leaky interzonal seals in stratified reservoirs can
allow fluid ‘leaks’ into overlying strata.

2.12.6 Reservoir Heterogeneities


Reservoir heterogeneities impacting in situ combustion recovery performance include permeability barriers to
lateral and vertical flow, natural fractures, high permeability thief zones, directional permeability, presence of gas
cap and aquifers.
Permeability barriers can have both positive and negative effect upon the in situ combustion process. As a positive
effect, vertical permeability barriers can divide a thick reservoir into smaller units, which may be more compatible
with the in situ combustion process. Vertical barriers can also act as a seal to upward migration of injected air and
may result in a more uniform burning in relatively thick reservoirs. As a negative effect, horizontal permeability
barriers can reduce the reservoir continuity and recovery.
Fractures and joints are secondary properties that may create preferential flow channels and influence the
recovery. A thin zone of high permeability at the top of the reservoir extending from one well to another
constitutes a hazard to fireflood by thieving air and starving the fire front of needed oxygen.
Directional permeability resulting from the anisotropy characteristics of the reservoir has a major influence upon

32
the performance of many in situ combustion projects. These include the Iola fireflood in Kansas (Hardy and
Raiford, 1975), Fry in situ combustion project in Illinois (Earlougher et al, 1970), and the Webster fireflood, in the
Midway Sunset field, California (Soustek, et al, 1994). Grain size and its orientation contribute to the existence of
directional permeability in a heterogeneous reservoir. Often the orientation of the medium to coarse-grained
sands establishes the direction of high permeability zone. Directional permeability can cause air to flow more
freely in one direction than in any other direction and result in uneven burn. Existence of directional permeability
alone is not a sufficient cause to reject a site for fireflood. By selectively locating the wells in the direction of
permeability, the recovery can be maximized.
Presence of free gas caps or thin layer of high gas saturation at the top of the sand is not a desirable geological
feature for fireflood operation because they can act as a thief zone for injected air and promote uneven burning.
Presence of bottom water leg or aquifer, though, not desirable from the point of anisotropy is not an impediment
to the success of a fireflood project. Many successful fireflood projects as the Glen Hummel, Gloriana, Trix-Liz, N.
government Wells fireflood in Texas were implemented in reservoirs with active aquifers. In these projects the
aquifer not only provided pressure support to the reservoir but also acted as a conduit to transfer the heat ahead
of combustion front.
Though, reservoir heterogeneity can have adverse effect on project performance their impact can be minimized
through recognition of the distinctive architecture of the reservoir and tailoring the project to accommodate this
architecture. Fire-floods implemented in highly heterogeneous reservoirs also often require unique reservoir
management strategies to make the project economically viable. Some examples of successful fireflood projects
where the combination of unique engineering design and reservoir management strategies overcame the many
conditions considered adverse to the success of the process include Unocal’s Brea-Olinda fireflood in Orange
county, California (Showalter, 1974), Mobil’s Moco fireflood in the Midway-Sunset field (Curtis, 1989; Soustek,
1991), and Mobil’s North Government Wells fireflood in South Texas (Casey, 1971).

2.12.7 Rock Properties


The key rock properties of interest to an engineer evaluating a prospect for the application of in situ combustion
process is sand texture, permeability and its distribution, porosity, and composition of rock matrix. In many
firefloods, especially those implemented in light oil reservoirs rock composition is more important than oil
properties in determining the amount of fuel available for combustion.

2.12.8 Sand Uniformity and Texture


Oil sands often vary considerably in their characteristics both vertically and laterally. The degree to which
heterogeneous sand approaches homogeneity or uniformity, however, impact the fireflood performance. Actual
grain size and grading, shape of grains, character and amount of cementing material determines the physical
characteristics and properties of the reservoir. The size, shape and sorting of the grains determine the porosity
and permeability of sand. Coarse, well-sorted and rounded sand grains result in a high porosity, high permeability
reservoir.
The permeability profile (permeability variation) as determined from core analysis, is valuable information for
determining the relative homogeneity of the sand. Generally the
greater the degree of uniformity exhibited in a profile more uniform the burn will be. However, there are many
instances of economically successful firefloods in sand bodies with relatively poor permeability profiles (Casey,
1971; Soustek, 1994). In the Mobil’s Webster reservoir combustion project, where sand bodies are lenticular and
anisotropic, the success of the project was attributed to tailoring the operating policy to suit the reservoir
architecture. In this reservoir the burn front advanced more rapidly through the high permeability medium to
coarse-grained sands than through finer-grained, thin to medium bedded sands, reflecting the influence of
anisotropy on the lateral rate of movement of burn. Proper reservoir surveillance and modification of production
strategy based on the timely identification of heat breakthroughs were cited as the keys to the success of the
project (Soustek, 1994).

33
Thus from an oil recovery and sweep efficiency aspect the degree to which a particular profile correlates from one
well to another is more important than the exact shape or dimensions of that particular profile.

2.12.9 Permeability
The actual value of permeability has very little effect on the mechanics of combustion process. Economically
successful firefloods have been implemented in less than 10 millidarcy carbonate light oil reservoirs (Miller, 1995).
The only requirement for permeability is that it must be adequate to permit air injection at a pressure compatible
with overburden at an acceptable compression cost. In viscous heavy oil reservoirs too low permeability may fail
to provide the minimum air flux needed for sustained combustion. Low permeability also increases air injection
pressure requirements and compression costs, and prolongs the operation. Low permeability in a viscous (greater
than 100 cp.) shallow reservoir can limit the injectivity and promote low temperature oxidation. In such reservoirs
a permeability greater than 100 millidarcies would be necessary.

2.12.10 Porosity
High porosity is desirable, since it directly reflects the volume of hydrocarbons that the rock can hold. In the U.S.,
economically successful firefloods have been implemented in reservoirs whose average porosity range from a low
of 0.16 to high of 0.38. As porosity decreases, the amount of heat stored in the rock increases. A lower porosity
will not have a significant impact on overall energy utilization in wet combustion process because part of the heat
stored in the burned volume of the reservoir will be recovered through scavenging operations. The main impact
of porosity will be in its oil content. The economic success of a fireflood is dependent more on the actual value of
the oil saturation-porosity product (fSo) than on porosity. Porosity lower than 0.2 is acceptable only if the oil
saturation is greater than 0.45.

2.12.11 Oil Saturation


Minimum oil content (the product of oil saturation and porosity) is necessary in order to offset the consumption
of oil as fuel in an in situ combustion process. A widely accepted rule-of-thumb in the industry is that if fSo is less
than 0.09 or 700 bbl/ac-ft. dry combustion should be eliminated from further consideration. This arbitrary cutoff
simply implies that the reservoir should have enough recoverable oil to cover the energy requirements of the
process and supply additional production to make the process economically attractive. For wet combustion where
the fuel lay down is lower somewhat lower oil saturation is acceptable.

2.12.12 Composition of Reservoir Matrix

The economics and applicability of fireflood in a reservoir is dictated to a large extent by the nature and amount
of fuel formed in the reservoir. If sufficient fuel is not deposited the combustion front will not be self-sustaining.
Conversely if excessive fuel is deposited, the process may result in dismal economics due to high air requirements,
high power cost and low oil recovery rate. Considerable laboratory and some field evidence exists indicating that
the mineralogical composition of the reservoir rock and chemical composition of the crude oil can effect the
amount of fuel available to sustain combustion.
Actual laboratory measurements of fuel formed using reservoir rock and crude indicate rock type is probably more
important than the crude properties, particularly in light oil reservoirs in determining the amount of fuel
deposition in the reservoir (Earlougher et al, 1970). The clay and metallic content of the rock, as well its surface
area has a profound influence on fuel deposition rate and its oxidation. Presence of clays and fine sands in the
matrix favor increased rates of fuel formation. Increase clay content particularly kaolonite and illite favor
increased rates of fuel formation by favoring low temperature oxidation reaction. Rock minerals such as pyrite,

34
calcite, and siderite also favor fuel-forming reactions. The catalytic activity of a reservoir mineral is highly
dependent on the specific composition of the crude. Low air fluxes resulting from reservoir hetrogeneneities and
oxygen channeling also promote low temperature oxidation and fuel formation reaction.
Results from the Fry in situ combustion project showed fuel deposition varied with the lithologic characteristics
of the rock. In the laboratory tests a very fine to fine grained sandstone containing significant amount of pyrite
and siderite deposited a greater amount of the fuel than medium grained sandstone containing similar amount
of pyrite. Similarly medium grained sandstone core containing large amount of clay material yielded the largest
amount of fuel (3.3 lb./ cu.ft. rock).

2.12.13 Effect of Well Spacing


Problems may arise in two ways when determining well spacing. If the well spacing is too close, the combustion
front may experience early gas breakthrough, while if the well spacing is too large, the oil production rate will be
slow, thus prolonging the life of the project and making the economic unattractive. Hence, the well spacing should
be in the optimum range to maximize oil recovery.
Geological considerations are quite important in determining the optimum flood pattern and well spacing. The
wells should be spaced to fit the geological pattern of the sand. Many of the sand bodies in the U.S. where fireflood
had been implemented are not continuous sheets of sand, but are lenticular in shape. These sand bodies
frequently exhibit anisotropy parallel to the bedding. The permeability of the sand in one direction therefore is
greater than in another. In such anisotropic, lenticular reservoirs, it is advisable from a stratigraphic standpoint
drill injection wells at right angle to the direction of high permeability trend and at closer spacing. The production
wells can be drilled along the trend on a wider spacing (600 ft. or greater). Such a flood pattern had been adopted
in some of the fireflood projects, implemented in the narrow “shoestring” pools of S.E. Kansas.

2.13 Injection of Oxygen-Enriched Air or Pure Oxygen

The performance of in situ combustion can be improved by injecting, instead of air, oxygen-enriched air or pure
oxygen. Moore received a U.S. patent on this method in 1965, but only recently has interest in its use increased
and several companies now have lab experiments and projects in operation (Cady and Moss, 1981; White and
Fairfield, 1982; Moore and Bennion, 1987; Petit, 1987). Evident advantages of the use of oxygen instead of air are:
- The decrease in the amount of gas to be compressed and injected because large volumes of nitrogen are
now eliminated. This increases the possibility of using in situ combustion in low permeability reservoirs and
reduces the volume of vent gases.
- Much higher solubility of C02 in the oil and water phases with oil viscosity reduction and significant swelling
of the oil. This increases the oil mobility and recovery and reduces the time to complete the project.
- The produced gas, rich in carbon dioxide, can be used as a source for other EOR projects.

However, when using oxygen as injected agent, strict safety precautions are required to prevent the formation of
a flammable mixture of oxygen-hydrocarbon gas (Burger and Sourieau, 1985).

The oxygen supply and injection system must be free of hydrocarbon material and not be oxygen reactive. In the
reservoir any unreacted oxygen breakthrough should be prevented by shutting in production wells. The down-
hole and surface equipment is exposed to highly corrosive conditions due to the high partial pressure of C02
produced together with a hot water phase.

35
2.14 Limitations of Combustion Process

Like all other oil recovery methods, combustion process has its share of shortcomings. Most of these limitations
can be overcome at considerable expense. This has reduced its overall attractiveness. Following are some less
attractive features of ISC:

- Though air is free, it must be compressed and delivered to the formation. The power required for
compressing air together with maintenance costs of the compressor are high enough that overall costs for
delivering air to the reservoir can be substantial. Relative to energy intensive steam injection operation, the
costs for in-situ combustion are lower only when the formation is less than 40 ft. in thickness. For thicker
reservoirs, the heat losses during a steam drive are low enough to enable the heat to be delivered at a lower
cost.
- Operational problems associated with combustion are more troublesome and require a higher degree of
technical sophistication to solve it. In comparison, steam injection operations are relatively problem free.
- Unlike the steam injection process design of in-situ combustion processes must be preceded by expensive
laboratory investigations. This is needed to ascertain the burning characteristics of the crude, fuel
availability and air requirements. Thus, planning and design of a combustion project is more expensive.
- Success and failure of combustion process implemented in reservoirs of similar and widely differing
characteristics preclude the development of satisfactory guidelines to screen reservoir for combustion
application. Expensive pilot is the only satisfactory answer to judge the feasibility of a particular project.
- The complexity of the in-situ combustion process hinder the development of more sophisticated numerical
simulators for complete performance prediction.

While considerable improvements are being made in the application of this technology, many operators still view
this technology as a high-risk operation. The commercial success of this process in the deep, extremely low
permeability carbonate, and clastic reservoirs in the U.S. had made operators take a second look at this process.
The success of horizontal well combustion technology in the heavy oil fields of Canada have also contributed to
revival of operators' interest in this process. Currently several new combustion projects are on the drawing board,
and one operator contemplates on implementing this process in a deep offshore light oil reservoir.
It is likely that the coming decade may see important advances in the application of this process in reservoirs
found in hostile environments. It is also likely that the process will increasingly be applied to recover light oil in
the U.S. and elsewhere. Thus future potential for oil recovery by ISC is very promising.

2.15 Design Considerations

Conditions favoring the use of ISC rather than steam include the following: 1) high reservoir pressure where steam
is not efficient, 2) potential for severe well bore heat losses (i.e., depth, offshore, permafrost), 3) reservoir clay
swelling in contact with fresh water, 4) limited water supply and 5) environmental regulations prohibiting steam
generation.
Like any other injection process, the design of ISC projects must consider injection pressure limitations and
reservoir flow resistance. These are especially important in heavy oil reservoirs where combustion must occur in
the high temperature regime to be successful. The minimum air flux needed to maintain high temperatures at the
front is estimated to be 0.125 ft/day (0.04 m/day). As the burn zone growth is directly proportional to the injected
air, the maximum air injection rate determines the minimum lifetime of the project. Ways to increase the air
injection rate are often needed, especially in heavy oil reservoirs. They may include reduced well spacing, cyclic
steaming of injectors and producers and an increase in injection pressure. These factors will determine the
compressor pressure and volume output.
There has often been some controversy over whether ISC projects should be developed using patterns or line
drives. Many early projects were started as pilots with a single injector. Usually this resulted in an inverted five-
spot pattern. These pilots behaved contrary to plan with the combustion front moving in only one direction
because of permeability variations, gravity effects, well spacing differences or a combination of these factors.

36
Attempts to correct the unbalanced flow included stimulating unresponsive wells and limiting withdrawal rates
of wells that produced excessive volumes of combustion gas. Generally these efforts did not have the desired
effect.
In retrospect, this reservoir behavior makes sense. Once a combustion front is even slightly asymmetric, the higher
temperature and thus higher mobility will cause greater flow in that direction. Thus the flow will become more
asymmetric, finally resulting in flow principally in only one direction.
Since it is often difficult to decide, a priori, which direction the major flow will take, operating plans should remain
flexible until field performance indicates what injection scheme best utilizes the flow directions.
For the above reasons many of the more successful ISC projects have been line drive operations that start near
the top of the reservoir and move down dip. In such an operation, the direction of the fire front is known. The
operating engineers can then plan their completion and operating history in a rational way that will mirror the
front movement and breakthrough history. This operating practice can be seen in most of the successful ISC field
projects that has been discussed before.

2.16 Use of In-Situ Combustion

In this section, the use of in-situ combustion as primary, secondary, and tertiary recovery processes is described.
Most of the in-situ combustion field projects have been performed for secondary oil recovery purposes, whereas
in some cases, combustion process was also employed as primary and tertiary recovery processes. Several
examples are given which explain why in-situ combustion was selected as a primary, secondary, or tertiary
recovery process for recovering oil at various depletion stages of a reservoir.
2.16.1 Use as a primary recovery process
In a review article, Poettmann (1964) stated that the sooner the combustion process is applied to a reservoir, the
better the results. The application of the in-situ combustion to a virgin reservoir is just one step ahead in the right
direction. It should be emphasized, however, that economics is the only criterion which determines the optimal
timing of starting not only an in-situ combustion but also other enhanced oil recovery processes.
In a Moco zone reservoir of the Midway Sunset Field, Kern County, California, evaluation of several operating
methods indicated that the heavy crude oil would be recovered most economically if the combustion process was
applied in the beginning of the project. Accordingly, the combustion was started in November 1959, whereas the
air injection begun in January 1960, at which time the cumulative oil production from the reservoir was 0.4% of
the oil-in-place. The reservoir pressure was only slightly lower than the initial value (1000 psi) of the virgin
reservoir. In 1971, the total recovery of the heavy crude was 9 million bbl or 24% of the original oil-in-place and
amount of air injected was 26 Bcf. The volume of the burned oil was estimated at 7.5% of the reservoir in 1971,
and was mostly confined to the area at the top of the structure around the injectors and nearby producers. Based
on the results of the South Belridge thermal recovery experiment (Gates and Ramey, 1958).

2.16.2 Use as a secondary recovery process


Several field projects were chosen to employ combustion as a secondary recovery process to provide a possibly
more efficient alternative method to waterflooding under certain reservoir conditions. In the Brea-Olinda Field,
California, the combustion project was started by the Union Oil Company, because two other secondary recovery
methods (waterflooding and cyclic steam processes) had failed (Showalter and Maclean, 1974). The waterflooding
was unsuccessful because of the extreme reservoir heterogeneity in which water tended to channel through the
formation excessively. Although the cyclic steam injection did increase crude oil production to some extent, the
added recovery was not economically beneficial compared with the costs.
A combustion project in a thin reservoir containing high-gravity oil in the May-Libby reservoir, Louisiana, was
operated by the Sun Oil Company. A detailed analysis of waterflood project showed that nearly 53% of the original
oil-in-place would remain in the reservoir at the end of waterflooding. An efficient recovery method was needed,
therefore, to recover the rest of the oil. Based on the analysis of laboratory and reservoir modeling results, it was
decided to initiate a fireflood in the western portion of the reservoir, while continuing waterflood in the eastern
37
portion of the reservoir.
In the Sun Production Company's Glen Hummel Field, Texas (Buchwald et al., 1973), waterflood, polymer flood
and in-situ combustion were evaluated. Among these three processes, combustion was finally selected based on
the expected recovery, profits, and anticipated life of the project.
In the Amoco Production Company's Shannon pool. Salt Creek Field, Wyoming (Parrish et al., 1962), primary
production method recovered only about 2% of the original oil-in-place in several decades. An analysis of the
reservoir indicated that oil recovery by waterflooding may not exceed 30% of the original oil-in-place. In addition,
several research results showed that this reservoir was favorable for in-situ combustion process. Finally, it was
decided to start in-situ combustion to recover oil from this reservoir.
A research group of the Continental Oil Company ( Martin et al., 1972) discussed in detail why in-situ combustion
was selected as a secondary recovery method in the North Tisdale Field, Wyoming. Water cuts had increased to
50% by the time the original development was completed. The primary recovery production declined to nearly
60% of its peak rate by 1957, and it was estimated that economic recovery by primary method would be less than
5% of the original oil-in-place. It was determined that conventional waterflood could yield additional recovery;
however, the presence of bottom water, the relatively high oil viscosity, and high water cuts were not favorable
for such a process. Moreover, other reservoir characteristics appeared to be favorable for fireflood, including
porosity, permeability, depth, and pay thickness. Among several variables, the main concern was the presence of
the hot water zone. Thus, the combustion process was selected to be employed for oil displacement.

2.16.3 Use as a tertiary recovery process


In several reservoirs, a considerable amount of oil still remains after successful completion of primary and
secondary recovery processes. A combustion pilot project was started when a successful waterflood was
approaching its economic limit in the Fosterton Northwest Unit, Fosterton Field, Saskatchewan, Canada (Marberry
and Bhatia, 1974). When the combustion project was started, the cumulative oil recovery by primary and
secondary methods was 23% of the original oil-in-place. The main aim of this combustion pilot study was to
evaluate the feasibility of employing in-situ combustion as a tertiary oil recovery process. If this pilot project is
successful, combustion process could be utilized in several reservoirs in southwestern Saskatchewan which are
similar in characteristics to those of the Fosterton Northwest Unit. Inasmuch as the oil viscosity is as low as 14 cP
at the reservoir temperature, other EOR methods were also considered. It was decided that in-situ combustion
was only one of the passible methods to enhance productivity of the waterflooded reservoir.
Three in-situ combustion pilots were conducted in the Potter Sand, Midway Sunset Field, California (Counihan,
1977). One of the three projects was unique and combustion was used as a tertiary oil recovery method after
steamflooding process. The main aim of this pilot project was to determine its feasibility and to compare the
results obtained from steamflooding and in-situ combustion flooding processes under similar conditions.
The combustion process was also employed as a tertiary recovery method in the Sloss Field, Nebraska, by Amoco.
The results of several studies, however, indicated that the reservoir was not an ideal candidate for in-situ
combustion. Reservoir properties such as permeability and porosity are reasonably good, but not outstanding.
The reservoir is thin and deep, containing light oil with low viscosity. In addition, the residual oil saturation at the
end of waterflood was low. A pilot project, however, was started. Later it was expanded to a full-scale project to
evaluate the economic success of employing combined forward combustion and waterflooding processes under
these reservoir conditions (Parrish et al., 1974).
These examples illustrate that the in-situ combustion process can be applied to recover the residual oil after
completion of waterflooding or other secondary methods.

2.17 Economic evaluation

Wilson and Root (1966) made an economic study in which they used a modified form of Chu’s two-dimensional
model. Their main objective was to compare cost of heating a reservoir either by steam injection or by forward
38
combustion, without regard to recover. Their sole consideration was the cost of heating to the same radial
distance by either forward combustion or by steam injection. They concluded that:

- Except for oils which yield coke in amounts less than 1 lb/cu ft and reservoir thickness of about 10 ft or less,
heating the reservoir by steam injection is cheaper than by forward combustion as long as the cost of the
fuel needed to generate the steam is low.
- For a given thickness, pressure and rate of heat injection, either process may be cheaper, depending upon
the reservoir fuel consumption and depth. However, as the price of fuel increases, the cost of heating by
steam injection increases more rapidly.
- Increased coke deposition favors steam injection.
- Increased wellbore losses by increased depth favors combustion.
- As the heated distance in the reservoir increase, heating by combustion becomes more favorable.
- As the sand thickness decreases and pressure increases, combustion is favored over steam injection.
- As the injection rates decreases, costs of steam injection becomes more favorable relative to air.

2.18 Disadvantages of in-situ combustion

There are several disadvantages of in-situ combustion, which can be summarized as follows:

1. The major portion of heat generated by in-situ combustion is not utilized in heating the oil. Instead, it heats
the oil-bearing strata, base and cap rocks, and interbedded shales. In-situ combustion, therefore, would be
economically feasible in those reservoirs which have high oil saturation and high porosity, as well as moderate
sand thickness (with less rock formation to be heated).
2. Several studies indicate that the reservoirs with viscous and low-gravity crudes are best suited for in-situ
combustion, because they provide the needed fuel for combustion. In addition, the required air/oil ratio for
viscous crudes is high, whereas their prices are, in general, lower than the high-gravity crudes. Some projects
for higher-gravity and lower-viscosity crudes, however, have also been reported to be successful.
3. The fireflood process has a tendency to displace oil only from top portion of the oil zone. The vertical sweep
in very thick reservoir formation, therefore, is likely to be poor.
4. There are several serious operational problems associated with in-situ combustion process. For example,
formation of emulsions causes pumping problems and reduces well productivity. The production of acidic
(low pH) hot water with high content of iron and sulfate ions causes environmental pollution and well
corrosion problems, whereas increased sand production and carvings cause plugging of well liner. In addition,
there is a possibility of (a) plugging of the producing wellbore due to deposition of carbon and wax as a result
of thermal cracking of the oil, and (b) liner and tubing failure due to excessive temperatures atl the production
well.
5. There is a large investment required for the installation of in-situ combustion equipment. Surface installation,
however, needs less fuel than hot-water or steam-generating equipment.
6. There is a possibility of explosion in the compressor. In order to minimize explosion hazards in the air injection
system, an explosion-proof lubricant should be employed.

2.19 Current Status of In-Situ Combustion

The in-situ combustion process is attractive economically, provided it is applied to petroleum reservoirs containing
approximately 50% oil saturation. The fuel content is one of the important parameters for combustion support at
a relatively low air/oil ratio. Laboratory studies on fuel content and air requirements can establish the possibility
of starting a field project. Although laboratory experiments can provide some basic understanding of the process,
the primary evaluation factor is a field application before the process is employed on a large scale.

39
The present status of oil production by in-situ combustion in the United States is nearly 11,000 bbl/day, whereas
the world oil production is approximately 58,000 bbl/day. It is likely that very little laboratory research can be
performed to improve the displacement efficiency of this process. With continued improvement of the in-situ
combustion technology, it is almost certain that some form of this process, such as dry, wet, and partially
quenched combustion, will find greater application in the coming years.

2.20 Case Study

Cambay basin of Gujarat is one among the promising petroliferous basins of India. The northern part of this basin
has a belt of heavy oil fields. Santhal field form a part of this heavy oil belt with crude oil viscosity ranging from 50
to 200 cps at an average reservoir pressure and temperature of 100 kg/cm2 and 70 °C respectively at 990m MSL.
The primary recovery factor as low as 13% has been envisaged in view of high mobility ratio between oil and
water. This necessitated implementation of an enhanced oil recovery technique. Based on extensive laboratory
studies, in-situ combustion process was found to be the most suited enhanced oil recovery method. For bench
marking the process parameters through field application an In-Situ Pilot in a similar heavy oil reservoir of adjacent
Balol field was initiated in 1990 and subsequently commercialized in 1997. Based on pilot results, the ISC process
was executed in KS-I reservoir adopting an inverted 5 spot injection- production pattern in the northern part of
Santhal field. However during commercial application the injection well pattern was modified to up-dip line drive.

In-situ combustion in Santhal field, primarily involves conversion of an existing oil producers to air injectors by re-
completion with heat shield , burner assembly and thermocouple run over tubing for temperature monitoring
at the surface. After initiation of chemical method of ignition at sand face, air injection in pre- estimated quantity
is maintained to sustain the burning of fire front. The ISC process is very much effective as only a small amount of
in-place oil is consumed while the rest is displaced, banked and finally produced. This process has a combined
effect in terms of efficient displacement of crude oil, reservoir pressurization with the formation of secondary gas
cap resulting in added advantage of gravity drainage, flue gas stripping of the reservoir oil, and oil swelling. In in-
situ condition, the injected air is utilized during combustion where two types of oxidation reactions namely LTO
and HTO occur which in turn generate heat energy. It helps in lowering the viscosity of oil and hence improves its
mobility. Ahead of the high temperature combustion front there exists a LTO reaction zone forming coke which is
utilized as a fuel for combustion front to sustain.

2.20.1 Reservoir Fluid Property

Santhal field has an anticlinal structure trending North-South direction. The reservoirs are sandstone, dipping 3 to
5 degree from west to east. There are five co-relatable pay sands designated as USP, KS-I, KS-II, KS-III and Lower
stack. well developed and hydro dynamically connected. Reservoir facies are seen pinching out up dip in the western
part abetting against Mehsana horst. The reservoirs have permeability of the order of 3-5 darcies, average porosity
of 28% and are operating under active edge water drive. Observed the reservoir oil viscosity increases from south
to north i.e from 50 to 200 cp. The crude oil contains 9-9.5% asphaltene, 10-13% resins. The density of the crude is
0.95 gms/cc at 15 °C. Targeted OIIP considered for ISC application is 41.4 MMt. Expected oil recovery by the
application of in-situ combustion process is 14.93 MMt from Santhal field.

2.20.2 Implementation Scheme

Prior to ignition of an air injector the base-line data of oil, gas, water quality and production testing parameters of
40
3the surrounding producers was generated. This is essential for comparison of parameters generated from pre and
post combustion process to evaluate the effect as shown in Table-I. The estimation of EOR gain is also being made
on the basis of improvement over the base line production data.
Ignition operation generates a fire front in the reservoir through the injector. This has been executed by artificial as
well as spontaneous ignition method. In both the cases the well bore is thoroughly cleaned and made free from
traces of oil to avoid explosion leading to tubling and casing damage. Detergents and hot water followed by Nitrox
solution are used for this purpose.
The artificial ignition allows generation of additional heat to create the fire front. The heat is generated in the well
bore by artificial means viz. gas burner, electric heaters etc. In case with Santhal ISC, a gas burner is tied with
thermocouple to monitor the ignition at the surface. TEB, an explosive chemical is being used for artificial ignition.
During ignition operation using gas burner, temperature within wellbore is maintained in a range of 350-450 0C
depending upon the casing grade by regulating air-gas ratio through measuring equipment attached to the ignition
trailor. The ignition operation is completed by giving heat load of approximately 1 MM Btu/ft to the formation.
However spontaneous ignition has also been achieved by simply injecting air into the reservoir and allowing the oil
to generate heat by oxidation and thereby ignite spontaneously. The spontaneous auto ignition process depends
on the oxidation potential of the crude and this parameter is determined in laboratory at reservoir conditions.
Once ignition takes place the well is put on regular air injection at an initial rate of 10,000 NM3/d. The air injection
rate is then gradually increased in steps to the peak level as determined by the reservoir volume to be heated. The
dry combustion phase continues at peak air rate for a minimum period of 90 to 100 days.
In order to utilize the enthalpy of the sediments behind the advancing combustion front and improving the sweep
efficiency, the well is subjected to wet phase combustion. It involves injection of pre-estimated volume of air in a
cycle of six days followed by one day water.

2.20.3 Process Monitoring

During the in-situ combustion process in Santhal field, following parameters are evaluated for monitoring purpose.
1. Daily record of Injected volume of air and water along with well head injection pressures.
2. Compositional analysis of produced gas by Gas Chromatograph using porapak column, ORSAT Apparatus and
Gas detector tubes.
3. Produced Oil and formation water quality.
4. Production test data: liquid rate, water cut, gas rate, well head temperature, tubing head pressure and casing
head pressure.
5. Reservoir pressure and temperature.
6. Corrosion study.

The compositional analysis data of produced gases indicate the nitrogen percentage has increased gradually up to
78% as compared to its base level of 1- 3% indicating the migration pattern of flue gases. It also provides an estimate
of the volume of the reservoir influenced under the effect of the ISC process. Similarly, ORSAT analysis of the gas
samples indicates that carbon dioxide percentage has increased gradually up to 20% from it’s initial level of 2-5%
and at the same time, oxygen concentration is observed to be less than 1% as compared to 20% in the injected air.
This signifies efficient utilization of oxygen in the combustion process. The year wise migration pattern of nitrogen
and carbon dioxide indicates that up-dip wells in the field were affected initially followed by wells in down-dip. It
was also observed that the nitrogen percentage is more (50-70%) in the up-dip wells compare to its value in the
down-dip (10-40%).Contray to increase in concentration of nitrogen in the up dip producers, increase in the carbon
41
dioxide percentage (8- 12%) in the down-dip wells was observed. This is possibly due to solubility of carbon di-oxide
in formation water within the reservoir which gets released at surface from the produced fluids.

The Acid number of crude oil gives an idea of the mode of oxidation reaction taking place during the process. The
higher acid numbers indicate low temperature oxidation whereas the lower acid number is due to high temperature
oxidation. The produced carbon mono- oxide and carbon dioxide also act as an indicator for knowing the mode of
oxidation of the process. Level of carbon mono oxide and carbon dioxide are also linked with LTO and HTO mode.
The increasing trend of carbon dioxide (15-20%) indicates high temperature oxidation occuring in the field. The
detailed analysis of viscosity, acid number, density of the crude oil and the chemical composition of the produced
water provides added information regarding the efficacy of the process and mode of combustion.

During the ISC process, oil viscosity had decreased in some locates in the field due to thermal effect. Decreasing
trend of oil viscosity is also supported by low acid values. Contrary to above, oil viscosity has been noticed to increase
in some area under the ISC influence. This is attributed to insufficient air injection, resulting in low temperature
oxidation. This also leads to less carbon dioxide generation which may in turn extracts light to intermediate
components from the oil leaving behind the heavier fractions in the reservoirs.

Hydrogen sulphide is often associated in the flue gases. The organo-sulphur compounds present in heavy oil and
pyrite present in the formation are the possible source of generation of hydrogen sulphide on their thermal
decomposition. It has been observed that the Hydrogen sulphide concentration in produced gases in Santhal field
varies between 100-1500 ppm. Higher concentration of Hydrogen sulphide in some of the wells is being neutralized
by injecting the caustic soda solution.

Online corrosion monitoring is done by introducing the coupon of the same material in flow line and periodically
measuring the weight loss after a fixed time of exposure as per NACE standard. The corrosion rate is used to select
the corrosion inhibitor and its dosage.

Presence of nitrogen and carbon dioxide in the produced gases has also been observed even the wells completed
in KS-III and Lower Stack sands underlying the KS-I and KS-II pay sands. This suggests vertical continuity since the
shale isolating these layers is stochastic in nature added information regarding the efficacy of the process and mode
of combustion.

During the ISC process, oil viscosity had decreased in some locates in the field due to thermal effect. Decreasing
trend of oil viscosity is also supported by low acid values. Contrary to above, oil viscosity has been noticed to increase
in some area under the ISC influence. This is attributed to insufficient air injection, resulting in low temperature
oxidation. This also leads to less carbon dioxide generation which may in turn extracts light to intermediate
components from the oil leaving behind the heavier fractions in the reservoirs.

Hydrogen sulphide is often associated in the flue gases. The organo-sulphur compounds present in heavy oil and
pyrite present in the formation are the possible source of generation of hydrogen sulphide on their thermal
decomposition. It has been observed that the Hydrogen sulphide concentration in produced gases in Santhal field
varies between 100-1500 ppm. Higher concentration of Hydrogen sulphide in some of the wells is being neutralized
by injecting the caustic soda solution.

Online corrosion monitoring is done by introducing the coupon of the same material in flow line and periodically
measuring the weight loss after a fixed time of exposure as per NACE standard. The corrosion rate is used to select
the corrosion inhibitor and its dosage.

Presence of nitrogen and carbon dioxide in the produced gases has also been observed even the wells completed
in KS-III and Lower Stack sands underlying the KS-I and KS-II pay sands. This suggests vertical continuity since the
shale isolating these layers is stochastic in nature.
42
2.20.4 Air Injection Pressure

So far 23 injectors are subjected to air and water injection in Santhal field. The conversion of producers to injectors
has been done in phase at different times to get the maximum recovery of oil. The well head Injection pressure with
time shows that the air injection pressure had an increasing tendency during initial phase which had later decreased
with the initiation of wet phase. This phenomenon may be attributed to formation of coke during the combustion
process that would have washed away during the wet phase. In general the injection pressure varies between 82-
100 Kg/cm2 at the well head.

2.20.5 Production Performance

The effect of ISC process is visible from the flow behavior of the producing wells. A number of wells which
were operating under artificial mode prior to in-situ combustion process have started flowing on self. It has
also been observed that wells located in the strike direction and up-dip was affected first followed by down-
dip wells. Due to gravity effect some of the up-dip wells have lately started producing only flue gases
requiring their forced closure. The up dip migration of flue gas is seen to be synchronized with gradual
increase in liquid production and decrease in water cut resulting increase in net oil production. The AOR
below 1000 Nm3/m3 indicates the ignition operation well within the economic limit which however shall
further decrease with time as oil production increases.. Overall performance of Santhal field shows that with
increasing numbers of injectors there is a gradual increase in oil production with time.

2.21 Summary – Conclusion

In-situ combustion is applicable to a wide array of reservoirs. In fact, it is the only thermal method that can
presently be applied to deep reservoirs, though deep downhole steam generation is being tested. It can be
used at any stage of reservoir depletion; it can be used in special situations such as offshore or in Arctic
regions. Because of the lack of heat losses at the surface and in the injection wells, it is the most thermally
efficient thermal recovery method. The injectant (air) is readily available. Combustion allows wider well
spacing than steam; economic results are comparable to those of steam injection.

Several aspects of operating in-situ combustion projects must be considered:


1. The large compression ratio and associated costs required to inject air into the formation
2. The planning and design requirements for a combustion project, which are more difficult than for
steam injection
3. Extensive laboratory work to assess fuel availability, air requirements, and burning characteristics
of the crude that are required before designing in-situ combustion projects
4. The high degree of technical sophistication and the careful monitoring needed to ensure proper
operation of a project
5. The limitation of numerical simulation and other techniques that makes predictions of recovery
more difficult than most other enhanced oil recovery methods.

Considerable improvements in the application of in-situ combustion have been made since the early
projects. New developments, such as application to light oil reservoirs, and the use of horizontal wells are
reviving interest in in-situ combustion. This process deserves consideration for many reservoirs, including
those in hostile environments or those not amenable to other recovery methods.
43
Reference

 F. John Fayers, (1981), “Enhanced Oil Recovery”.


 W.Fred Ramirez, (1987), “Application of Optimal Control Theory to Enhanced Oil Recovery -Elsevier
Science Ltd”.
 Erle C. Donaldson_ G. V. Chilingarian, (1985), “Enhanced Oil Recovery I_ Fundamentals and Analyses
(Developments in Petroleum Science) (v. 1)”.
 Erle C. Donaldson_ G. V. Chilingarian, (1989), “Enhanced Oil Recovery_ II_ Processes and Operations
(Developments in Petroleum Science) (v. 2)”.
 N. K. Baibakov_ A. R. Garushev_ W. J. Cieslericz, (1989), “Thermal Methods of Petroleum Production
(Developments in Petroleum Science)”.
 Aurel Carcoana, (1992), “Applied Enhanced Oil Recovery”.
 Partha S. Sarathi, (1998), “In Situ Combustion Handbook – Principle and Practices”.
 Partha S. Sarathi, (1998), “Nine Decades of Combustion Oil Recovery — A Review of
 Combustion History and Assessment of Geologic Environments on Project Outcome”.
 S.Stevens, V. Kuuskraa, J. O’Donnell, (1999), “Enhanced Oil Recovery Scoping Study”.
 Teknica Petroleum Service Ltd, (2001), “Enhanced Oil Recovery”.
 M. Nabipour, (2007), “An Introduction to Enhanced Oil Recovery”.
 Laura Romero-Zerón, (2012), “Introduction to Enhanced Oil Recovery (EOR) Processes and
Bioremediation of Oil-Contaminated Sites”.
 E. Tzimas (2005) ‘’Enhanced Oil Recovery using Carbon Dioxide in the European Energy System.
 New Billions in Oil
 SPE-89451-MS

44

You might also like