You are on page 1of 16

SPE-179536-MS

The Myths of Waterfloods, EOR Floods and How to Optimize Real Injection
Schemes
Richard Baker, BRE Group, Adj. Professor U of C; Rous Dieva, Baker Hughes; Robert Jobling, and Crystal Lok,
Baker Hughes

Copyright 2016, Society of Petroleum Engineers


This paper was prepared for presentation at the SPE Improved Oil Recovery Conference held in Tulsa, Oklahoma, USA, 1113 April 2016.
This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents
of the paper have not been reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect
any position of the Society of Petroleum Engineers, its officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written
consent of the Society of Petroleum Engineers is prohibited. Permission to reproduce in print is restricted to an abstract of not more than 300 words; illustrations may
not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract
Waterflood implementation accounts for more than half of the oil production worldwide. Despite the
observations and extensive research from a large number of floods and thousands of simulation studies,
managing waterfloods and Enhanced Oil Recovery (EOR) floods is still a technical challenge. A major
contributor to this challenge are waterflood induced fractures (WIF). Managing waterfloods is a multivariable problem although WIF are one aspect, it is by no means the only controlling factor.
The best evidence that WIF are one of the main factors controlling flow in reservoirs is the insensitivity
of injection pressure to injection rates. With our experience, in hundreds of waterfloods, we have
frequently observed this phenomenon in the field data. If fluid flow depended on diffusive Darcy flow
alone, we would expect higher injection rates with higher injection pressures. However, it is common to
observed relatively constant injection pressures over a wide range of water injection rates. Rapid well
communication and changes in water cuts that vary with injection rates also support an interpretation of
high permeability induced fractures between injector and producer. In some reservoirs, interwell tracer
data can be used to determine the influence of induced fracture features. The interwell tracers usually
show very fast water movement.
Induced fractures in waterfloods and EOR projects can be caused by a number of mechanisms such as
but not limited to, pressure depletion, changing pressure regimes, thermal effects, or plugging effects.
These fractures can either be beneficial to the reservoir performance or effect performance negatively.
Benefits include improved injectivity and increased throughput of the displacing fluid. Negative effects
can come in the form of reduced volumetric sweep efficiency, impaired ultimate recovery or injected fluid
losses out of zone.
Case studies, theory, and available literature from Western Canada will be reviewed in order to suggest
and improve reservoir management strategies for waterfloods. We have completed hundreds of waterflood
feasibility, waterflood management and EOR flood studies worldwide and continue to be amazed and
humbled by the complexity that many waterfloods and EOR floods exhibit due to induced fracturing. WIF
and EOR induced fractures (EIF) are common and should be analysed to optimize production. Growth of
the WIF, response to waterflood with the presence of WIF, implication of WIF and reservoir management
are the main areas which will be addressed.

SPE-179536-MS

What are the Myths of Waterflooding?


1. All flow in the reservoir is dominated by matrix properties only

We cannot assume that all flow in the reservoir is dominated by matrix properties alone. This
is true even in reservoirs that are not thought to be naturally fractured reservoirs.

2. Excellent characterization of matrix properties alone guarantees accurate forecasts

Excellent characterization of matrix properties alone will not guarantee accurate forecasts
because of flow through WIF.

3. Long induced fractures cannot and do not develop in the most reservoirs

Long fractures can develop and cause premature water breakthrough especially in situations with
high injection rates, cold water injection, dirty water injection, or low permeability reservoirs.

4. Reservoirs with long induced fractures cannot achieve high recoveries

Even in reservoirs with long fractures high recovery factors are possible by smart reservoir
management through; monitoring, injector-producer conversions, and pattern reconfiguration.

5. Permeability in the reservoir is typically modelled only as a function of relative permeability and
saturation. Therefore, changes in time due to pressure and stress effects are not normally accounted for

Effective permeability especially around injectors cannot be treated as a constant; effective


permeability can vary with rates.

6. As waterflooding has been a well-developed and a well-used recovery method, a detailed daily
monitoring program for the injectors is not needed

Compared to a conventional waterflood, a waterflood with large propagating fractures has the
added uncertainty of the architecture of the fractures. There is a critical need to monitor
injection rates and wellhead pressures daily and have frequent production well tests to help
partially characterize the fracture systems.

7. Recovery factors of waterflooding and other injection schemes are thought to be rate insensitive

Some reservoirs with waterflooding or gas injection have recovery factors that are rate sensitive.

8. In most waterfloods, it is assumed that injection pressure increases with injection rate

Wellhead fracture extension pressures are subtle therefore monthly injection pressure may
seem constant regardless of the rate.

Introduction
The following phenomena are commonly observed from the presence of waterflood induced fractures (WIF);
Many waterfloods have relatively long fractures caused by thermal effects, pressure depletion (poroelastic) or plugging effects1-13. These fractures may be beneficial in that they help injectivity and injection
rates. However long fractures greater than inter-well distances can hurt recovery in some producers.14
These WIF can have long lengths but fortunately height growth is limited. This is based on our
experience with production logging of vertical injectors. Temperature, radioactive and other production
logging tools show relatively good confinement of injected fluids into the pay zone. Operational strategy
can also play a role in injected fluid confinement.

SPE-179536-MS

These WIF are not homogenous, planar-like fractures with infinite conductivity but instead in many
reservoirs the WIF seem to be complex as shown in Figure 1 below. There are substantial pressure drops
in some of these complex fracture systems.

Figure 1Fracture system types

1. These WIF have dimensions that are injection rate and injection pressure sensitive. Fractures may
close if injection rates and pressures are reduced.
2. In systems with long fractures relative to the inter-well distance, leak off from fractures to matrix
dominates long term performance. Imbibition and cross flow from matrix to the fracture can be the
determining factor in a waterfloods ultimate recovery.
3. In systems with long fractures relative to the inter-well distance short term water breakthrough is
often controlled by WIF fracture growth and fracture length.
4. Matrix permeability, precipitations of waxes17, particulate matter and water injection temperature
can in some cases strongly affect fracture length and thus waterflood performance.
All of these phenomena can be divided up into three major categories which will be discussed: the
growth and direction of the WIF, waterflood response where WIF are present, and reservoir management
of such floods. This paper focuses on the WIF but in EOR or gas flooding projects some of the same
phenomena are present and observed.
A step back will be taken from research on induced fractures some global observations will be made.
The mental model that the industry holds about waterfloods or other injected floods is partially obsolete
due to the presence and influence of WIF, especially in low permeability reservoirs (10 mD). The mental
model most of us carry is that matrix permeability alone controls volumetric sweep. From our experience,
a rough estimate would be that 60% of waterflood reservoirs worldwide are affected by WIF.

Growth of Waterflood Induced Fractures


Injection of low viscosity fluid such as water or gas being the cause of the WIF is very different from
conventional high viscosity fracture fluids used in a hydraulically fracture treatment. This is due to high

SPE-179536-MS

fracture leakoff rates in water injection or gas cases. In large hydraulic fracture treatments often gel and
fluids inhibit leakoff.
Waterflood or injection induced fractures can be a very slow process in comparison to hydraulically
induced fractures. WIF growth periods can take days, months or years depending on the reservoir
(permeability, stresses, etc) and the pressure regime. Simulation studies of injection induced fractures
show rapid growth initially but growth slows down after ~10 to100 days*1.
The Spraberry CO2 pilot showed the slow WIF growth18. Spraberry is a field with low matrix
permeability (1 md) that is a naturally fractured sandstone reservoir. During the CO2 pilot there was an
initial re-pressurization of the pilot area by water injection, followed by CO2 injection. During the initial
re-pressurization by water injection four separate interference tests were conducted.
Water injection was started at four injectors and pressure was monitored at the offset producer wells.
The pressure response at offset producers clearly showed that there was no fracture initially from injector
to producer; there was no initial deviation from the pressure trends, but over a period of 8-10 days the
fracture grew between some injectors and some producers. There was marked deviation from the pressure
trend at this stage. The distance between injectors and producer is ~600 ft; however, when the injectors
got shut in pressure decrease at the offset producers was immediate. This response indicates an open
fracture and was observed at four separate interference tests between water injectors and offset producers.
The slow transition from delayed to instantaneous pressure response leads us to conclude that fracture
growth in the Spraberry field, off trend from the natural fractures is not instantaneous.
Fracture growth is strongly dependent on fluid leakoff from the fracture to the matrix, therefore matrix
permeability is a critical controlling factor of WIF growth. Damage in the form of reduced communication
between the fracture and the matrix can be a significant concern. The introduction of particulates or wax
precipitation17 will reduce leak off rates from the fractures which can cause a negative feedback loop,
causing the fracture to increase in length and potentially cause earlier breakthrough. Damage due to
particulates, waxes, etc, cause a reduction in leak off rate from the fractures to matrix and therefore may
lead to longer fracture growth8, 16, 19.
Near wellbore reservoir temperature is another controlling factor for WIF growth. A decrease in
temperature due to the injection of cold water relative to the original formation temperature can cause a
contraction of the rock leading to a lower fracture extension pressure1, 3-11, 16, 18 and increase the potential
for WIF growth. For most reservoirs worldwide bottom hole injection temperatures are cooler than
formation temperatures.
Another WIF growth phenomena is the potential to connect existing natural fractures over time or
simply connect areas of weakness over time. Water injection can widen existing fracture systems and
connect otherwise discontinuous natural fractures. This is critical to reservoir management as reservoirs
that show this type of behavior will often have significant variability in water breakthrough behaviour.
As discussed above fluid leakoff from the fracture to the matrix is the key variable in fracture growth.
If the rate if injection into the fracture exceeds the leakoff from the fracture to the matrix than there is
significant risk of fracture growth. In this scenario the pressure in the fracture would increase until it
reached the fracture pressure of the rock and the fracture would continue to grow until the injection and
leakoff rate balanced again.
WIF will continue to grow if the local voidage replacement ratio is greater than 1.016. This fracture
growth profile has important practical implications; often operators will inject at a voidage replacement
ratio (VRR) greater than 1 initially in order to quickly re-pressurize the reservoir. High VRR may result
in fracture extension. It is recommended that instantaneous voidage replacement ratio stay below 1.5 to
*1 The time to stabilize the fracture length depends upon many variables such as matrix permeability, temperature, stress, reservoir pressure, fracture damage
and voidage replacement ratio therefore this is a rough general estimate.

SPE-179536-MS

decrease the chance of long fractures. However, the injection pressure and near wellbore pressure
compared to the fracture pressure of the reservoir rock must be monitored.
In Figure 2, an idealized view of fracture growth is shown. WIF, as seen by the dotted black ellipses,
are compared to a conventional hydraulic fracture treatment, as seen with the yellow elipse. Note the t2,
t3 may take days or months to reach that effective length due to high leakoff from WIF, while in
conventional hydraulic fracture treatments, length is typically reached in the time frame of hours.

Figure 2Time dependant fracture growth

Analysis of Injection Pressures and Rates in WIF Reservoirs


Figure 3 clearly shows the lack of sensitivity of injection rate versu injection pressure in the Wainwright
field, which is a heavy oil, sandstone waterflood. Note that this injector even at later stages of waterflood,
has minimal changes in reservoir pressure and near wellbore saturations and thus transient effects have
dissipated and static reservoir pressure is constant. Also note that large increases in injection rate did not
result in increases in injection pressure.

Figure 3Injection pressure in red and water injection rate in green.

SPE-179536-MS

Another observation of water injection rates and pressure in certain injectors is their cyclical nature.
When we observe frequent injection rates and corresponding wellhead injection pressures we often see a
rapid increase in injection pressure with constant injection rates followed by a rapid decline in pressure.
Figure 4 shows this cyclic phenomena for a low permeability waterflood within a sandstone rock. Note
for this plot the operator was trying to maintain constant rates and pressure. This cyclic phenomena does
not appear when the data is taken less frequently say at monthly intervals. This observation has been seen
in other parts of the world as described by other authors.19-22

Figure 4 Injection pressure and water injection rate on a tight low permeability water injector

The same relationship can also be seen with increasing pressure and increasing injection rate. Based
on this relationship, the phenomena of the WIF can be explained using Darcys Law seen below.

If injectivity is proportional to injection rate, assuming the viscosity, geometry factor and drawdown
pressure is the same, then the effective permeability has to be changing to incorporate those increasing
injection rates. Therefore, it is observed that the system is adjusting with time. Overall, it is important to
understand that permeability is not a static parameter in most waterfloods due to the WIF and their
dynamic nature.
Other authors have also noted this constant injection pressure with variable rates23 or changes in
permeability with changes in injection rate25.
WIF can often exhibit a very subtle fracture pressure as we can see from Figure 3 and 5. Note that
the bulk of the injection pressures are clustered at 4,000 kPa in Figure 3. Monthly injection pressures and

SPE-179536-MS

injection rates do not seem to correlate. This is most likely due to the negative feedback loop exhibited
during fracture extension as mentioned above. Other literature also demonstrates such phenomena29.

Figure 5Inter well tracer test30

Fracture Length and Complexity


WIF systems are complex fractures and their length can vary, as seen in Figure 1. The observations that
support complex fractures especially in both the on trend and off trend direction is the high water retention
(70%) in the initial stages of waterflooding (~first 3-4 years). If simple linear fracture systems existed
between wells we would expect water to travel along these fractures and the affected wells to water out
rapidly. We very often see an initial delay in water production even in wells that strongly communicate
in later stages of the waterflood showing fracture complexity and imbibition. When capillary forces are
strong and imbibition is strong, water gets imbibed into the matrix slowing down the initial water
breakthrough.
We see this fracture complexity when examining inter-well tracer tests in mature waterfloods. Figure
5 below shows a tracer test in the Wainwright field. The producing wells are ~400 m away from the
injectors. We can see rapid breakthrough of inter-well tracers at on-trend (NE-SW) orientated wells in 4
days and 26 days. What is also interesting is that the fall off production profile (after peak tracer

SPE-179536-MS

concentration) does not show a rapid decline as you would expect in a simple, single, linear fracture
system instead we see a more gradual (diffuse) tracer production response in the fall off period. This tracer
profile suggests a complex fracture system that has multiple branches.
As observed by Wright26, in West Texas using tiltmeter data, the half lengths of the WIF can be 70%
to 100% of the well spacing as seen in Figure 6. The next important aspect in understanding the growth
of a WIF that needs to be analyzed is the direction of that growth.

Figure 6 WIF half length26

Direction of Waterflood Induced Fractures


Generally, WIF seem to follow the major stress direction which in Western Canada is in the NE-SW
direction. This trend seems to fit the majority of Canadian waterfloods and gas floods data24,25. However,
not all fractures, including induced fractures, appear to be exclusively in the NE-SW trend.1,26 Please refer
to Figure 7 which shows the natural stress trends in Western Canada and the United States based mainly
on borehole breakouts. In our experience, in both the USA and Canada we see these stress trends affecting
waterfloods/EOR production. However, it is more common to see east-west communication in the USA
as you would expect from Figure 7.

Figure 7Stress Regime of Western North America

SPE-179536-MS

A very interesting study was performed by Penn West in Alberta, Canada27. The example presented In
Figure 8 is from a Cardium sandstone pilot with low matrix permeability. These wells were initially
fractured by hydraulically induced fractures. The performance of two horizontal well pairs drilled in
different orientation, as seen in the figure below, were tested and analyzed. The major stress trend is in
the NE-SW direction and will be referred to as the on-trend.

Figure 8 Cardium Well pair 1 drilled off-trend and Well pair 2 drilled on-trend to the regional fracture orientation.

Figure 9 shows Well Pair 1 and figure 10 shows well pair 2 performance. Displayed in the top graph
are oil rate in green, water injected rate in light blue and liquid production rate in black; the bottom graph
shows watercut in dark blue and GOR in red. Well Pair 1 was drilled NW-SE and both wells were
hydraulically fractured. Observing the data, the fracture network most likely grew along the major stress
direction NE-SW (perpendicular to the wellbore). When observing the production data, after start of
injection end of 2009, no response in oil production is seen. However, a sharp increase in watercut is seen
at the beginning of injection, suggesting connectivity between injector and producer and potential WIF
growth.

10

SPE-179536-MS

Figure 9 Well Pair 1 Performance

SPE-179536-MS

11

Figure 10 Well Pair 2 performance

However, when water injection rate was decreased the watercut decreases correspondingly which
shows that once a WIF is established with different operating practices the connection may be changed.
Tilt meter and micro seismic tests were also completed by Penn West. Figure 11 shows Well Pair 1 at
the bottom left corner and Well Pair 2 at the top right corner along with the tilt meter and microseismic
results. With well Pair 1 it can been seen that the tilt meter shows fractures growth perpendicular to the
well bore, which supports the direction of the fracture network growth in the NE-SW direction, or the on
trend direction. In this case, water communication is observed faster, as seen from the watercut profile in
Figure 10, which increased from ~65% to ~85% within months of water injection.

12

SPE-179536-MS

Figure 11Tilt meter and microseismic test performed27

Figure 10 shows well Pair 2 which was drilled SW-NE, along the major stress direction, and it was also
hydraulically fractured. In this case, the fracture network would grow in the direction of the major stress
or parallel to the wellbore. In this particular case, observing the production data it can be seen that
watercut increased gradually from 10% to ~40% in over 4 years. The oil rate is seen to stabilize and
increase when liquid increase in the last 2 years of production. A more stable flood front is observed in
this example.
When the tilt meter and micro seismic data is analyzed as seen in Figure 11, it can be seen that it
supports fracture growth in the NE-SW direction. There is noise in the tilt meter data, suggesting that some
of the fractures in the producer (10-28) may be in the off-trend direction. However, the majority of the
micro seismic data is seen clustered along the wellbore, which is in the NE-SW major stress direction.
Overall this example demonstrates that knowing the local reservoir pressure, the initial stress orientation,
and choosing appropriate operating practices can significantly change the performance of an area. Once
the fracture network is established, changes in injection rate and static reservoir pressure can potentially
close some of those fractures and change the performance of the area.
Our final field example shows two behaviors that we have mentioned above. First, is that the WIF
grows at a slow rate in most cases. Second, is that WIF can grow in an off-trend direction.
Induced fractures, as mentioned earlier, are complex and can also grow in the off-trend direction. An
example of a Cardium sandstone, tighter oil field in Alberta29, will be presented below. An image of the
well configuration, which includes both vertical and horizontal wells, is shown in Figure 12.

SPE-179536-MS

13

Figure 12Cardium field

In this particular example, the waterflood induced fracture growth is observed in the off-trend direction.
The focus of the example is the well pair in the North part of the field highlighted in the red ellipse. The
associated production data is displayed in Figure 13.

Figure 13North Well Pair Production response

The well pair displayed shows a steep watercut increase once injection begins. Watercut and water
injection rates are displayed, along with the static reservoir pressures measured at the injector and

14

SPE-179536-MS

producer. Note, there is an initial delay in the response to water injection. For two years the water cut at
the off-trend producer is low. The bottom right had corner of Figure 13 shows the static reservoir pressure
reading taken at the injector (16-27). Note the rise in reservoir pressure in 1979-1980. After that we see
a rapid rise in water cut for producer 100/02-34. It is important to note that the induced fracture direction
is in the NW-SE which is in the off trend direction.
This example shows WIF growth in the off trend fracture direction. The rapid rise in watercut cannot
be explained by any other phenomena. The fact that it took a number of years before this significant
watercut increase was seen supports the relatively slow growth of WIF discussed in previous sections.
This example proves that care needs to be taken in understanding the reservoir and knowing the injection
pressure and injection water rate relationship when determining potential fluid flow pathways. This
relationship is critical to appropriate reservoir management of WIF floods.

Reservoir Management
It is critical to understand that WIF can have both positive and negative effects. Most of the waterfloods
referred to in this paper are successful waterfloods with recoveries greater that 30% OOIP. Many floods
may see only beneficial effects of WIF in that injectors can have better injectivity especially if they are
in lower permeability reservoirs (10md).
Often in land based operations, producer-injector conversions and changing injection rates cause
dramatic improvements in recoveries. Imbibition of water from the fractures to the matrix mitigate some
fracture effects, this can be a critical force in low permeability waterfloods. We often see the initial
positive response in the direction of the fracture, however at later stages the off trend wells respond better
as they do not have water breakthrough as quickly. Waterflood recovery is rate sensitive in some
reservoirs, therefore using Darcys law and fractional flow theory the higher the injection rate the better
the economics, however induced fractures mean that too high an injection rate will cause short circuits and
water cycling resulting in poor sweep.
The combination of positive and negative results means there is an optimum injection rate that
maximizes recovery.
The good news is that due to stress effects even, long WIF are not irreversible. We have seen cases where
decreases in injection rates have made dramatic decreases in water cuts with low lag time. However, with time
and more crossflow high water saturation channels may develop and make the effects of WIF more permanent.
Because of water crossflow from the fractures to the matrix and generally the matrix having a higher water
relative permeability, even if the fractures close a preferential flow path may still exist.
Good reservoir management entails 1) reducing injection rate at some injectors, 2) doing producer to
injector conversions to maintain voidage replacement ratio. Gel treatments can also be effective in some
areas with small fracture volumes but encased in consolidated rocks.

Conlcusions
In conclusion, after reviewing the existing myths of waterflooding, a need is found for a revised workflow
when dealing with waterflood induced fracture systems. Managing waterflood is a multivariable problem
although WIF are one aspect, it is by no means the only controlling factor. Below are a number of
observations and analysis which should be included when evaluating WIF floods and further understanding the reservoir.
1. We cannot assume that all flow in the reservoir is dominated by matrix properties alone. This is
true even in reservoirs that are not thought to be naturally fractured reservoirs. When examining
waterflood reservoirs offset wells and other wells outside of the nominal area should be analyzed
to get a better reservoir understanding. It is important to understand dynamic injection rates and
pressures.

SPE-179536-MS

15

2. The best evidence that WIF are a main controlling factor of flow in reservoirs is the insensitivity
of injection pressure to injection rates. This is often observed in field data. If we had diffusive
Darcy flow alone, we would expect higher injection rates which would yield higher injection
pressures but often injection pressure is observed to be constant. Rapid well communication as
seen by pressure behavior, production well tests and water cut profiles that vary with injection
rates can also suggest WIF.
3. Excellent characterization of matrix properties alone will not guarantee accurate forecasts because
of flow through WIF. Crossflow between the WIF and the matrix as well as the established
preferential flow pathways are critical.
4. Long fractures can develop and cause premature water breakthrough especially in situations with
high water injection, cold water injection, dirty water injection or low permeability reservoirs.
Analyzing the production/injection, pressure data and water cuts for the wells of interest may help
to determine the areas with potential long fracture growth.
5. Effective permeability especially around injectors cannot be treated as a constant; effective
permeability can vary with rates.
6. Compared to a conventional waterflood, a waterflood with large propagating fractures has the
added uncertainty of the fractures architecture. There is a critical need to monitor injection rates
and wellhead pressures daily and have frequent production well tests to help partially characterize
the fracture systems.
7. Distance can be an optical illusion in optimizing waterfloods. Fluids can move large distances
relatively fast, especially if they are associated with faults or large permeability pathways. This
behavior is seen time and time again when pressure tests are conducted and wells further away see
the response.
8. Some reservoirs with waterflooding or gas injection have recovery factors that are rate sensitive.
9. Wellhead fracture extension pressures are subtle therefore monthly injection pressures may be
misleading and may seem constant regardless of the rate.
10. Fracture pressure is not a constant dependent on depth alone, temperature, pressure depletion and
stress regimes are critical in the understanding fracture behavior.

References
1. Richard Baker, Tim Stephenson, Crystal Lok, Predrag Radovic, Robert Jobling, Cameron McBurney; Analysis of
Flow and the Presence of Fractures and Hot Streaks in Waterflood Field Cases, SPE 161177, Calgary, October-2012.
2. Ojukwu, K.I., Van Den Hoek, P.J.: A new way to Diagnose injectivity decline during fractured water injection by
modifying conventional hall analysis, SPE 89376, Tulsa, April 2004.
3. Martins, J.P., Murray, L.R., Clifford, P.J., McLelland, W.G., Hanna, M.F., Sharp Jr., J.W., Produced water reinjection
and fracturing in Prudhoe Bay, SPE 28936, New Orleans, September 1995.
4. Ovens, J.E.V., Larsen, F.P., Cowie, D.R., Making sense of water injection fractures in the Dan Field, SPE 52669,
San Antonio, October 1997.
5. Perkins, T.K., Gonzalez, J.A., The effect of thermoelastic stresses on injection well fracturing, SPEJ (Feb 1985) 78,
AIME, 279.
6. Wright, M.S., Svendsen, A.P., Clifford, P.J., Thermally induced fractures of Ula Water injectors, SPE 20898, The
Hague, October 1990.
7. Andrews, J., Hettema, M., Nesse, T., Injection wells: A case study from the Statfjord Field, SPE 90949, Houston,
September 2004.
8. Gadde, P.B., Sharma, M.M., Growing injection well fractures and their impact on waterflood performance, SPE
71614, New Orleans, October 2001.
9. Griffin, L.G., Wright, C.A., Demetrius, S.L., Blackburn, B.D., Price, D.G., Identification and implications of induced
hydraulic fractures in waterfloods: Case history HGEU, SPE 59525, Midland, March 2000.
10. Hustedt, B., Zwarts, D., Bjoerndal, H.P., Masfry, R., Van den Hoek, P.J., Induced fracturing in reservoir simulations:
Application of a new coupled simulator to waterflooding field examples, SPE 102467, San Antonio, September 2006.

16

SPE-179536-MS

11. Van den Hoek, P.J., Dimensions and Degree of containment of waterflood-induced fractures from pressure-transient
analysis, SPE Reservoir Evaluation & Engineering, October 2005.
12. Van den Hoek, P.J., Sommerauer, G., Nnabuihe, L., Munro, D., Large-scale produced water re-injection under
fracturing conditions in Oman, SPE 87267, Abu Dhabi, October 2000.
13. Heffer, K., Fracture characterization through rate correlation analysis, SPE 154429, Copenhagen, June 2012.
14. Dyes A.B., Kemp C.E., Caudle B.H.; Effect of Fractures on Sweep-out Pattern; AIME TP 8033 volume 213, 1958.
15. Dikken, B.J., Niko, H.; Waterflood-Induced Fractures: a Simulation Study of their Propagation and Effects on
waterflood sweep efficiency, SPE 16551, Aberdeen, September 1987.
16. Settari, A., Warren, G.M.; Simulation and field analysis of waterflood induced fracturing, SPE 28081, Delft, August
1994.
17. King R.W. and Adegbesan K.O.; Resolution of the Principal Formation Damage Mechanisms causing Injectivity and
Productivity Impairment in the Pembina Cardium Reservoir, SPE 38870, San Antonio, October 1997.
18. Baker, R.O., Contreras R.A. and Sztukowski D.; Characterization of the Dynamic Fracture Transport Properties in
a Naturally Fractured Reservoir, SPE 59690, Midland, March 2000.
19. Slevinsky B.A.; A Model for Analysis of Injection-Well Thermal Fractures, SPE 77568, San Antonio, Oct. 2002.
20. Martins, J.P.,Murray, L.R., Clifford, P.J., McLelland, G.,Hanna, M.F.,Sharp, J.W. Jr.; Produced-water Reinjection
and Fracturing in Prudhoe Bay, SPE 28936, Alaska, August 1995.
21. Williams, D.B., Sherrard, D.W., Lin, C.Y., (1989), The Impact of Waterflood Management of Inducing Fractures in
Injection Wells at Prudhoe Bay, SPE 16358, Venture, April 1987.
22. Charlez, P., Lemonnier, P., Ruffet, C., Bouteca, M.J., Tan, C.,(1996), Thermally Induced Fracturing: Analysis of a
Field Case in North Sea, SPE 36916, Milan, October 1996.
23. Baker R. and Gregor V.; Fracture Characterization Of The Wainwright Pool; CIM paper 88-39-45 June 1988.
24. Morales R.H., A.S. Abou-Sayed and A.H. Jones; Detection of Formation Fracture in a Waterflooding Experiment,
SPE 13747, October 1986.
25. Heffer K. J., Fox R.J., McGiii C. A., and Koutsabeloulls N. C.; Novel Techniques Show Links between Reservoir
Flow Directionality, Earth Stress, Fault Structure and Geomechanical Changes in Mature Waterfloods., SPE 30711,
June 1997.
26. Wright C.A., Understanding Hydraulic Fracture Growth: Tricky but not Hopefess, 1999 SPE 56724, Houston,
October 1999.
27. Government of Alberta Energy Department, 05-088 HMS Fracturing for Improved Waterflood Recovery 2011
Annual Report, 2011 IETP Annual and Final Reports., http://www.energy.alberta.ca/3783.asp
28. Geologic Systems Ltd, GeoScout data base, Alberta, Canada.
29. Warren, G.M., Settari, A.; Practical Aspects of Pressure Induced Waterflood Fracturing, CIM 95-11, May 1995.
30. Baker, R., Gregor, V.; Fracture Characterization of the Wainwright Pool, PETSOC 88-39-45, Calgary, June 1988.

You might also like