Sie sind auf Seite 1von 30

SHAHJALAL UNIVERSITY

OF SCIENCE & TECHNOLOGY, SYLHET


Department of PME
ASSIGNMENT ON STRESS DEPENDENT

SUBMITTED
TO:
PERMEABILITY
FOR TIGHT GAS
RESERVOIR

SUBMITTED TO:
Mr Arifur Rahman
Assistant Professor
Dept. of PME, SUST

SUBMITTED BY:
Jumman Al Jawad
Reg No: 2010336013
4th Year 2nd semester
PME, SUST

Tight Gas Reservoir:


Tight gas is the term commonly used to refer to low permeability reservoirs that produce
mainly dry natural gas. Many of the low permeability reservoirs that have been developed in
the past are sandstone, but significant quantities of gas are also produced from low
permeability carbonates, shales and coal seams. Production of gas from coal seams is covered
in a separate chapter in this handbook. In this chapter, production of gas from tight sandstones
is the predominant theme. However, much of the same technology applies to tight carbonate
and to gas shale reservoirs. Characteristics of tight gas reservoir show :

Gas reservoirs: low permeability and tight, thin effective thickness of single layer,
dispersive longitudinal distribution
Strong heterogeneity of reservoirs and low single-well control reserves
Quick pressure drop and low single-well production
Short stable production period and low average single-well production

Tight gas reservoirs have one thing in common: a vertical well drilled and completed in the tight
gas reservoir must be successfully stimulated to produce at commercial gas flow rates and
produce commercial gas volumes. Normally, a large hydraulic fracture treatment is required to
produce gas economically. In some naturally fractured tight gas reservoirs, horizontal wells
and/or multilateral wells can be used to provide the stimulation required for commerciality.

Figure: A typical tight gas reservoir

To optimize the development of a tight gas reservoir, the geoscientists and engineers must
optimize the stress distribution, the number of wells drilled, as well as the drilling and
completion procedures for each well. Often, more data and more engineering manpower are
required to understand and develop tight gas reservoirs than are required for higher
permeability, conventional reservoirs. On an individual well basis, a well in a tight gas reservoir
will produce less gas over a longer period of time than one expects from a well completed in a
higher permeability, conventional reservoir.
Tight gas reservoirs can be:

Deep or shallow
High pressure or low pressure
High temperature or low temperature
Blanket or lenticular
Homogeneous or naturally fractured
Single layered or multilayered

The low permeability of tight gas reservoirs is attributed to the complex diagenetic history which formed
the present day fabric of the reservoir. Understanding these processes is critical to predicting production
response and develop well designs and completion strategies which exploit existing reservoir
characteristics. Tight gas sands are often deeply buried and subject to large compressive stresses,
altering the original depositional fabric of the rock.
Actually, Unconventional gas resources constitute a large component of the remaining gas resource in
North America. While unconventional resources do not play as large a role in the resource base in the
rest of the world, it is expected that they will play a larger role as understanding of these resources
improves and as operators become more proficient at deploying advanced technologies necessary for
economic development. As the conventional gas resource base matures and output declines,
unconventional resources will play a more prominent role in supplying a growing energy demand.

Characterization
One of the challenges in developing low permeability resources is the ability to accurately evaluate rock
properties (such as permeability, porosity and capillary pressure). This information is necessary to
quantify resource potential and predict production behavior; essential information required for
developing viable energy policies and making sound investment decisions for field development. Due to
the complex structure of the pore network, simple relationships relating permeability to porosity are not
representative. The low permeability of these rocks makes the standard steady-state techniques, widely
applicable in evaluating conventional gas reservoirs, inadequate and virtually impossible to implement.
Application of these techniques to low permeability rocks yield unrepresentative results. Conventional
methods for assessing in-situ reservoir properties (such as shut in pressure build up tests) are not
practical due to the very long times required for the reservoir to reach pressure equilibrium and the
delay of production (or income from the well) as a consequence of the shut in.

Stress Distribution:
Soils that effect of external load are subjected to stress. The vertical stress increases in soil due
to various type of loading. At any point in soil the stress applied from own weight of soil which
called effective stress, and from external load which called net stress, the net stress which
applied must be determined.
The stress on element:

As shown in figure, the stresses on element are as follow:


On plan

XY the stress are

z, zx ,zy

YZ

the stress are

x, xy, xz

ZX

the stress are

y, yz, yx

From this z, x, y its stresses called normal stresses, but the other is called shear stresses
which is xy, yx, zy, yz, xz, zx.

Permeability:
Permeability is the property of rocks that is an indication of the ability for fluids (gas or liquid)
to flow through rocks. High permeability will allow fluids to move rapidly through rocks.
Permeability is affected by the pressure in a rock. The unit of measure is called the darcy,
named after Henry Darcy (1803-1858). Sandstones may vary in permeability from less than one
to over 50,000 millidarcies (md). Permeabilities are more commonly in the range of tens to
hundreds of millidarcies. A rock with 25% porosity and a permeability of 1 md will not yield a
significant flow of water. Such tight rocks are usually artificially stimulated (fractured or
acidized) to create permeability and yield a flow.

Figure: Permeability differences of different types of reservoir

Stress dependent Permeability:


In reservoirs with a significant stress-dependent permeability, reservoir models should include
stress-dependent permeability to improve accuracy for purposes of oil and gas reserve
determination and reservoir modeling. The benefits include a better understanding of the
behavior of tight gas sands, lead to a more accurate modeling of that kind of unconventional
reservoirs and get a more realistic forecasting of production performance and well test analysis.
The rate of permeability decline with increasing net effective stress is different for each rock
type and is controlled by three interrelated, pore geometrical parameters, length, and shape
and short axis dimension of the throats13, 19. Others important parameters are clay content,

pore volume compressibility and authigenic cementation. The mechanisms of permeability


reduction are much more pronounced in tight formations3. It can be expected that formations
with pore distribution of smaller radio are very sensitive to compressive stress.
The dependence of permeability on pore pressure makes the flow equation strongly nonlinear.
To study fluid flow through stress-dependent porous media, a new parameter, permeability
modulus or , is defined by Nur et al. and studied by Pedrosa and Kikani et al. as follows:

This parameter plays a very important role in systems where changes in effective stress affect
permeability. Basically, it measures the dependence of hydraulic permeability on pore pressure.
For practical purposes, the permeability modulus is assumed constant. Thus, permeability
varies exponentially with pore pressure.
In view of the similar appearance of permeability and density in the diffusion equation, it may
be advantageous to assume an exponential relationship between permeability and pressure.
This choice has some experimental support and mathematical convenience shown by Kikani
and Pedrosa.

Stress dependent Permeability for tight gas reservoir:


Tight gas reservoirs are characterized by having poor rock properties. Those reservoirs typically
have low porosity and permeability. Tight gas reservoirs have been considered as gas storage
rock with low quality. A tight gas reservoir is generally recognized as any low permeability
formation which special well completion technique are required to stimulate production.
Typical values of porosity are lower than 10% and permeability is usually below 0.1 md.
There are some fundamental differences in rock-water-gas interactions between tight
sandstones and normal gas reservoirs. These differences result primarily from significant pore
structure alterations as the rock undergoes compaction and diagenesis.
As gas production begins from the reservoir, pore pressure decreases and the effective
stress increases; the relation between these variables is shown in the following equation:

where,
S corresponds to total stress,
is the effective stress (matrix stress, grain to grain pressure) and
p is the fluid pressure.

Equation states that every change in the pore-fluid pressure under otherwise constant
conditions, result automatically in a change of the effective stress. Rock permeability in tight
sands is significantly affected by changing the effective stress.
The behavior of tight gas sand permeability in response to changing effective stress can be
explained qualitatively by the complex and tortuous pore structure that results from extensive
compaction and diagenesis. Thin section and scanning electron microscope (SEM) images of the
pore structure reveal very narrow slit-like apertures between pores.
These thin cracks provide the major connectivity, which allows fluid to move when the rock is
under low effective pressure conditions. However, increasing effective pressure easily closes
such flats cracks.

A second order, nonlinear, partial differential equation results when variation of permeability
with pressure is considered in the continuity equation. A different kind of flow-reducing
mechanism has been studied experimentally by a number of investigators. This mechanism is
the reduction in permeability caused by an increase in effective frame stress.
In the reservoir an increase in effective frame stress is caused by fluid withdrawal and the
accompanying decrease in pore pressure. Since the overburden force on the reservoir rock
remains the same, the decreasing pore pressure results in an increased effective frame stress.
Because low permeability formations are more affected by stress changes, this effect can be
expected to be more significant in deep gas reservoirs.
When fluid flows through porous medium, not only the natures of the fluid affect the flow, but
also the character of the rocks and the interaction between the fluid and rocks. With the
extraction or injection of the fluid, the framework of the rocks may become deformed because
of the changes of the stress. The deformation would affect the physical properties of the
formation and the production of the oil fields.

Based on the studies, they propose several empirical formulas between permeability and
effective stress Most of the fluid-solid coupling formulas only take external stress into
consideration, and need improvement because of the lack of general applicability.

The impact of effective stress


Permeability decreases in step shape with the increasing of effective stress. But the decreasing
degree became slow down. As the core number 29 shown, the effective stress increased from
4.8 MPa to 6.7 MPa during the early loading stage. The increasing value of effective stress
occupies only 20% of the total changes, but the decreasing value of permeability occupy 62% of
the total changes. The permeability decreased obviously with the increasing of effective stress
owing to the pore throat with low closure pressure closing at the low effective stress stage. This
impact became less obviously because the quantities of the throat with easy closing characters
decreased at the later stage.
The permeability would not recover while the fluid pressure recovers. It proved that ductile
deformation which damaged the original flowing channel happened in the pore structure of the
rocks. The damages were not reversible.

Stress sensitivity coefficient of permeability


Rock permeability would change obviously because of the change of the stress field when fluid
flow through the low permeability reservoir. The fluid-solid coupling relation formula had no
actual physical significance at present. To solve this problem, our paper presented the stress
sensitivity coefficient of permeability just as the definition of compressibility coefficient of
porosity

where,

Ko = initial permeability,
K=permeability changes,
P=effective stress changes
Ck means unit permeability decreasing value at effective pressure increasing unit pressure. It
indicated the sensibility of permeability when pressure made change. The horizon was more
sensitive when Ck was greater.

EXPERIMENTAL PROCEDURE
All measurements were performed on cylindrical shaped rock samples of 1.5 in diameter and
1.0 in length. The samples were cleaned using toluene and alcohol in Soxhlet type extractor and
then dried in a vacuum oven.

Measurement of Porosity and Permeability


The porosity and permeability of the rock samples were measured simultaneously at every step
of the pressure cycle. A combined porosity and permeability measurement apparatus was
assembled for this purpose (Abdulraheem et al., 1999). The porosity was measured using the
Helium gas expansion method. The permeability was measured either by the steady state
method or by the pressure pulse decay technique (Figure 1) for very tight samples. As
mentioned earlier, the range of confining pressure was 0-82 MPa. The pore pressure during
porosity and permeability measurements is negligible compared to the externally applied
hydrostatic confining pressure. Nine simultaneous measurements of porosity and permeability
in this range during loading and four measurements during unloading were made. Enough time
was given at every pressure step for the stress to equilibrate and the corresponding strains to
fully develop. It is to be noted that only one method of measuring permeability is used for all
the pressure steps during a loading and unloading cycle. Pressure pulse method is used only for
a small number of sandstone samples with initial permeability less than 0.1 md at 4.12 MPa
confining pressure.
The procedure and theory for measuring the porosity by gas expansion method and
permeability by steady state method can be found in the standard text (Tiab and Donaldson,
1996). A brief review of determination of permeability by pressure pulse decay method is
provided below.
Pressure Pulse Decay Method
The schematic diagram showing the experimental setup of the transient pressure pulse decay
method is shown in Figure 1. The procedure can be described in the following points:
The system consisting of the core holder and the upper and lower reservoirs is brought to a
certain pressure called the system pressure.
The upper reservoir is isolated and its pressure is increased by about 2-3% of the system
pressure.
The pressure pulse is made to flow through the rock specimen and its decay with respect to
time is recorded by the data acquisition system. The pressure decay data can be used to
determine the permeability of the rock specimen.

Establishment of universal model


Quantitative characterization of pore structure

The stress sensitivity was affected by both the internal pore structure and external effective
pressure. Although pore structure is aberrant, it is statistical uniformity and self-similar. Fractal
theory is applicable to rocks. Pore fractal number could be calculated from capillary pressure
curves

where,
Pi means mercury injection pressure, Pa;
Vi means mercury injection volume;
rn means pore radius corresponding to mercury injection, um;
Vn means the total injected volume;
Df means pore fractal dimension number;
c is a constant.

Then we assume

Figure represents the capillary pressure curves of the four samples. Pore fractal number was
calculated by capillary pressure curves as figure 5. It can be seen that pore fractal number could
reflect the pore structure very well. But it costs a lot and needs large work capacity to test the
capillary pressure of every core. Fractal dimension number (Df) matches very well with horizon
quality factor ( K / ) in semi-log plots from the statistics of the fractal dimensions of thirty
actual cores as the next figure.
Form the research results above, it can be seen that horizon quality coefficient which
represents the horizon pore structure is an external symbol of pore fractal dimension number.
And the two parameters can be acquired easily.

Hydraulic fracturing design and tight gas reservoir simulation


A commercial simulator used to model the hydraulic fracture created during the stimulation
along with the stimulated natural fracture network using low-viscosity fluids. Stress profiles and
other elastic rock properties estimated in the geomechanical analysis were used as input for the
design. To achieve better proppant distribution, a low-viscosity linear gel was combined with
slickwater in the treatment. The low-viscosity linear gel was optimized using different
concentrations of ingredients for the high reservoir temperature (~126C) using source water
and local ingredients. Due to the high closure pressure and low viscosity of the fluid, highstrength small-mesh proppants were used in the design.

Effect of stress-sensitive coefficient on IPR curves


Productivity curves of fractured vertical gas well in different stress-sensitive coefficients are
plotted in Figure. From which we can see that as the stress-sensitive coefficients increase, well
productivity decreases and the decline rate of gas production is higher, meanwhile the
productivity curves bend over to the pressure axis in the earlier stage with a greater bending.
This is because the larger stress-sensitive coefficient indicates more intense of stress sensitive
effect to induce more pronounced decrease of permeability with the decline of reservoir
pressure, which impacts significantly on well productivity.

Damage Mechanisms in Tight Gas Reservoirs


Tight gas reservoirs can be subject to different damage mechanisms during well drilling,
completion, stimulation and fracturing, such as mechanical damage to formation rock, plugging
of natural fractures by invasion of mud solid particles, permeability reduction around wellbore
mainly as a result of filtrate invasion, clay swelling and liquid phase trapping .Materials such as
mud filtrate, cement slurry, or clay particles may enter the open pores of the formation and
reduce permeability around the wellbore as well. The solids may penetrate only a short
distance into the rock matrix and cause only a shallow mechanically damaged zone. However

the damage to natural fractures and open permeable conduits can be severe, as drilling fluids
invasion mostly occurs through the natural fractures. Liquid invasion damage into the rock
matrix is one of the major factors that cause low productivity in tight gas reservoirs. In the
absence of external cake protection, filtrate invasion into the tight formations is huge due to
the tremendous amount of capillary pressure suction that potentially imbibes and holds the
invaded water in the porous media.
The weak mud cake and strong capillary pressure suction may amplify the water invasion
profile and deteriorate severity of the phase trap damage to the tight formation.
The greater the difference between initial water saturation and critical water saturation in tight
formations the more significant is the potential damage to gas permeability.
A typical tight gas reservoir may produce mainly dry gas, and contain very low amounts of
heavy components. Therefore, condensate banking may not be a major issue in the tight gas
wells, and the well can be allowed to flow at low flowing bottom-hole pressure for higher gas
production rate. In addition, producing gas at low flowing bottom-hole pressure may reduce
the water phase trapping damage near the wellbore, since at lower pressure, water content of
the gas phase increases and water phase may partially be vaporized into the flowing gas phase
in the reduced pressure zone around the wellbore.
However in the cases that are not truly dry gas situation, producing with the large pressure
drawdown may cause condensate banking in the reservoir near the wellbore, if the flowing
bottom-hole pressure drops below the dew point pressure of the gas phase.

Oil based fluids may be considered in some situations for low permeability gas reservoirs. In the
case of oil-based drilling fluid invasion, there is no external water being introduced into the
formation and the fluid saturations and wettability remain unchanged. However invasion of the
oil filtrate into the tight formations may result in introduction of an immiscible liquid
hydrocarbon around wellbore, causing entrapment of an additional third phase in the porous
media. In the case of oil-based fluids invasion into water wet gas reservoirs, the invaded oil may
tend to be trapped in the central portion of the pore space, rather than adhering tightly to the
matrix walls as the wetting phase. Although this central pore space occlusion can cause
substantial reductions in permeability, in some cases, the damaging effect in overall is less than
the case where water based system is used in the same circumstances. Some types of oil may
also dissolve in the gas and clean up after some time. The relative permeability curves show the
reduced effective permeability due to the water invasion into the formation, compared with
damage to permeability caused by oil invasion.

Relative Permeability Effects (Phase Trapping)


This issue is one of the most severe that often plagues the success of a low permeability gas
reservoir production operation. Since many reservoirs of this type fall into the classification of
sub-normally saturated or desiccated reservoirs, there is a tremendous amount of potential
capillary pressure energy (capillary suction as it is sometimes referred to) which wants to
imbibe and hold a fluid saturation in the porous media. Phase trapping effects can occur in gas
reservoirs for both water and hydrocarbon based fluids. In most cases, water is the wetting
phase, which tends to reduce or eliminate the affinity for spontaneous imbibition of a
hydrocarbon based liquid phase into the matrix surrounding the wellbore (although this fluid
may still be displaced into the matrix of the rock by overbalanced drilling and completion
procedures, or deposited in place by the depressurization of rich retrograde condensate type
reservoir gases).
The basic mechanism of a phase trap in a low permeability gas saturated matrix is illustrated in
Fig. 8. It can be observed that the pore system of the reservoir is initially at a low liquid
saturation which provides the maximum cross sectional area for flow in the pore system, and
therefore the highest level of permeability.
If a water-based fluid is introduced into the system, we can see that a high water saturation in
the flushed zone is generated and results in some trapped gas saturation. Upon reversal of flow
to clean up the well, insufficient capillary drawdown gradient is present to overcome the
capillary pressure effects which results in a much higher trapped liquid saturation being
obtained in the porous media. The configuration of the gas-water relative permeability curves
for the porous media will determine the amount of reduction in permeability associated with
this increased saturation. Figure illustrates favorable and unfavorable rock geometries for
phase trapping issues with water.
It is observed that the pore system contains substantial microporosity and that the majority of
the effective permeability in this media is contained in a relatively small fraction of the pore

space which consists of interconnected meso or macropores or small fractures. In this case, the
natural capillary imbibition will draw invaded water into the tightest portions of the pore
system on a selective basis.
Although these portions of the reservoir can be saturated with water, effective gas permeability
may not be significantly altered as the occluded portion of the pore system represents solely
ineffective porosity. It is only when the trapped water saturation increases to the point where it
is sufficiently large that it begins to encroach into the meso/macropores and significant
reductions in permeability to gas are observed. Therefore, rocks of this pore geometry may be
significantly less sensitive to water-based phase trapping, as significant increases in the initial
"trapped" water saturation can be tolerated without severe accompanying reductions in
permeability.

Petro-physical Attributes of Low-permeability Reservoirs and


Implications for relative Permeability Mechanisms
The most significant differences between conventional reservoir and low-permeability
reservoirs lie in the low-permeability structure itself, the response to overburden stress, and
the impact that the low-permeability structure has on effective permeability relationships
under conditions of multiphase saturation. Figure provides a comparison of traditional reservoir
behavior with low-permeability reservoir behavior. In a traditional reservoir, there is relative
permeability in excess of 2% to one or both fluid phases across a wide range of water
saturation. Further, in traditional reservoir, critical water saturation and irreducible water
saturation occur at similar values of water saturation.
Under these conditions, the absence of widespread water production commonly implies that a
reservoir system is at, or near, irreducible water saturation. In low-permeability reservoir,
however, irreducible water saturation and critical water saturation can be dramatically
different. In traditional reservoir, there is a wide range of water saturations at which both
water and gas can flow. In low-permeability reservoir, there is a broad range of water
saturations in which neither gas nor water can flow. In some very low-permeability reservoir,
there is virtually no mobile water phase even at very high water saturations.
Because of the effective permeability structure of most low-permeability reservoir, there is a
large range of water saturations over which both water and gas are essentially immobile. A lack
of water production (or recovery from a test) should not be used to infer that the rocks are at,
or near, irreducible water saturation nor should these regions be regarded as water free.
Instead, low-permeability reservoir rocks should be regarded as having insufficient permeability
to either gas or water over a wide range of water saturations.

Figure highlights the relationships between capillary pressure, relative permeability, and
position within a trap, as represented by map and cross section views in conventional and low
permeability reservoirs. In both cases (A) and (B), the map illustrates a reservoir body that thins
and pinches out in a structurally updip direction. In conventional reservoir, water production
extends downdip to a free-water level (FWL). In the middle part of the reservoir, both gas and
water are produced, with water decreasing updip. The updip portion of the reservoir is
characterized by water-free production of gas.
In low permeability reservoirs, significant water production is restricted to very low structural
positions near the FWL. In many cases, the effective permeability to water is so low that there
is little to no fluid flow at or below the FWL. Above the FWL, a wide region of little to no fluid
flow exists. Farther updip, water-free gas production is found.

Based on the petro physical studies and the relative permeability variations in low permeability,
poor-quality reservoir rocks concluded that the gas fields in the Greater Green River basin are
not examples of basin center or continuous-type accumulations, nor are they a unique type of
petroleum system as generally believed. All these occur in conventional structural,
stratigraphic, or combination traps rather than regionally gas saturated unconventional basin
centered type.
Low-permeability reservoirs have unique petro-physical properties, and failure to fully
understand these attributes has led to a misunderstanding of fluid distributions in the
subsurface. An understanding of multiphase, effective permeability to gas as a function of both
varying water saturation and overburden stress is required to fully appreciate the controls on
gas-field distribution as well as the controls on individual well and reservoir performance.
A better understanding of the relationship between rock fabric and gas productivity requires
careful investigations into multiphase permeability under conditions of varying water saturation
and net-overburden stress, as well as an analysis of capillary pressure and net-overburden
stress.

GAS-WATER RELATIVE PERMEABILITY


The relative permeability in tight gas reservoirs has unique characteristics compared to that in
conventional reservoirs. The values of critical and irreducible water saturations in tight gas
reservoirs are not close to each other as is the case in conventional reservoirs. When the flow
of water is involved, relative permeabilities of all fluids in question are also important. Shanley
et al.20, presented a permeability jail concept based on relative permeabilities measured on
tight cores. As shown in Fig. 2, the critical water saturation (Swc) and the irreducible water
saturation (Swi), are not close to each other as is the case in conventional reservoirs. The
saturation interval between Swc and the Sgc could be a wideinterval, and the relative
permeabilities for gas and water are zero. It is important to determine the relative permeability
curves for a given reservoir at in-situ stress condition.

STRESS-DEPENDENT PERMEABILITIES
An experimental procedure was designed to simulate the reduction in reservoir permeability
(matrix, natural fractures and induced fractures) as a function of increasing effective stress.
Whole core samples were used with dimensions of 4 diameter and various lengths (4 - 8).
The sample was then tested for matrix conductivity. A natural shear fracture from a cored
formation was used to study the hydraulic conductivity of a shear fracture as a function of
effective stress. A proppant bed was sandwiched between the two split halves to form a
propped fracture. The sample (intact, with tensile fracture with natural shear fracture, or
propped fracture) was then positioned inside the rock mechanics loading frame. Then the
confining pressure was applied around the sample, and a linear flow was established at a given
pore pressure to determine the hydraulic conductivity of a given porous medium component.

STRESS-DEPENDENT MATRIX CONDUCTIVITY


Selected samples that did not appear to have microfractures were tested to determine the
stress-dependent matrix permeability. Various combinations of net effective stresses were
applied and the permeability was measured at each stress level. Table 1 presents all
combinations of applied confining stresses and the pore pressure levels, and gradients for a
given flow test.
Recalling the definition of effective stress as given in Equation, we must assume a value for
Biots coefficient (). Assuming is 1, and plotting matrix permeability as a function of effective
stress, we obtain Figure. Close examination of Figure suggests that for a given effective stress,
multiple values of permeabilities are measured. This is not an experimental error; rather the
assumption that can be 1 is not valid.
The next step is to change and replot the stress-dependent permeability function until a
meaningful trend is obtained. Since is a function of stress, then varying it within the
constraints from the first step would produce the stressdependent permeability presented in
Figure, with an estimated function (p).

STRESS-DEPENDENT TENSILE FRACTURE CONDUCTIVITY


When the reservoir pressure decreases, the elastic displacement in response to the increase in
effective stress will cause natural fractures to close, leading to a decline in reservoir
productivity. The matrix medium feeds the natural tensile fractures and the latter conducts the
fluids to the wellbore. The decline in conductivity with increasing effective stress should follow
a logical declining rate to support a given production rate. The elastic closure response occurs
when the net effective horizontal stress increases as a result of reservoir depletion.
The elastic response to close the fracture follows Hookes law of elasticity and it is controlled by
Youngs modulus of the formation:

The aperture of the fracture will decrease, causing a corresponding reduction of fracture
conductivity. If we assume 50 ft of the rock perpendicular to the fracture will contribute to

fracture closure, then for Youngs modulus of 3 x 106 psi, the decrease in fracture width
corresponding to a decrease in reservoir pressure from 7,000 to 4,000 psi will be 0.05. The
fracture will not close by 0.05, rather the contact points (asperities) will carry the applied
stress to prevent fracture closure if they are strong enough to withstand the stress. The
compressive strength of the asperities will determine the final fracture permeability. The
reduction in conductivity is due to a combined effect of elastic response and compressive
failure of the asperities. Compressive failure also generates rock particles and fines that will
further reduce fracture conductivity. The fracture asperities in tensile and shear fractures differ
considerably, as the former is not accompanied by formation shifting, while the rock in the
latter experiences formation shifting and so generates a higher conductivity. As explained
earlier, a 4 sample, shown in Fig. 5 failed in tension following a Brazilian test. The induced
failure plane represents a tensile fracture. The flow testing through the fracture was
performed, and results were separated from the total permeability using the following
equation:

STRESS-DEPENDENT PROPPANT HYDRAULICFRACTURE CONDUCTIVITY


Proppant hydraulic fracturing is used to create a conductive fracture in the pay zone to enhance
well productivity. Proppant is used to keep a fracture open during the life of the well.
A tensile-failed sample was propped with 100 mesh Intermediate Strength Proppant (ISP), or
with 30 mesh Intermediate Strength Resin Coated Proppant (ISRCP) and flow testing was
performed at variable effective confining stress.
The permeability of a one layer 30 mesh ISRCP decreased drastically, and a lot of fines were
generated, Figure, at an effective closure of 4,000 psi. This is an important criterion to consider
when deciding on the type of proppant to be used in the proppant fracturing treatment of a
given reservoir.
Figure shows the stress-dependent permeability of the porous media components considered
in this study. The best reservoir management will be produced if the conductivity of all
components decline at the same rate, assuming they start at similar original values. Figure
shows the stress-dependent conductivity for the same porous media components.
Normalized fracture conductivity, defined as a percentage of the initial conductivity, is
presented in Figure. A 100 mesh proppant fracture will sustain well productivity better than 30
mesh ISRCP, which was crushed at a closure stress of 6,500 psi effective stress. The 30 mesh
proppant conductivity is less than the ISP 100 mesh conductivity when the effective stress
exceeds 7,000 psi. Although this high stress may not occur deep in the reservoir, near the
wellbore it may be very applicable. Low concentrations of proppant (Waterfracs) have been
very successful in stimulating tight gas reservoirs. A Waterfrac is a proppant fracturing;
however, the proppant concentration used is very low, ranging from 0.5 lb/gal to 2 lb/gal.

Therefore, it is possible that a 100 mesh proppant will perform better than a 30 mesh resincoated proppant if the latter is exposed to crushing stress because fines will be generated from
the resin coat, reducing fracture conductivity below that of a 100 mesh proppant. Tensile
fractures lose conductivity at very early stages of reservoir depletion; therefore, it is anticipated
to have a sharp decline in production rate followed by a more stable rate for the life of the well.
Observing figure it can be concluded that filling natural fractures (tensile and shear) will
improve productivity of the given stress sensitive reservoir under study. This is not necessarily
true in non-stressed reservoirs, when the effective stress is less than 2,000 psi, the shear
fracture provides better conductivity than a 100 mesh ISP propped fracture.
Figure suggests that the tensile and shear fractures lose more than 70% of their conductivity
when the pore pressure decreases by 1,000 psi. This confirms the rapid decline in productivity
in naturally fractured reservoirs during early production. Propped fractures decline much more
slowly than natural fractures.

WATER BLOCK PHENOMENON/ UNDERBALANCED DRILLING

A formation damage mechanism called water block hinders the economic production rate from
tight gas reservoirs. Drilling and completion practices must be designed carefully for stress
dependent reservoirs. Upon drilling, a gas bearing formation may exhibit gas show. A
spontaneous imbibition is triggered, due to the high capillarity force of the tight formation.
Water imbibition continues until water saturation increases near the wellbore and/or the flow
is slowed down by mud cake development. The water phase flow into a gas bearing formation,
due to the capillary effect.
Initially, water saturation is at its irreducible value; thus, the capillary effect is higher than the
gas pressure gradient, causing water to continue getting imbibed until equilibrium occurs22.
During drilling, the gas pressure gradient drives gas from the formation into a wellbore, causing
a gas show experienced by drillers. At the same time, water is imbibed from the wellbore into
the formation. The imbibed water may become a water block zone, such that the pressure
drawdown is not enough to displace water from this zone. Underbalanced drilling does not
hinder water imbibition or water block development. Therefore, in tight gas wells, using oilbased fluids is far more effective than underbalanced drilling.

OPTIMIZING HYDRAULIC FRACTURES


To construct a commercial production in tight gas sand, a well must contact a large drainage
area as much as possible; therefore, wells drilled in these challenging reservoirs must be
hydraulically fractured. It is vital to understand the unique characteristics of tight gas reservoirs,
especially in the areas of formation damage and hydraulic fracturing.
Defining the porous components, especially the matrix/fracture interplay, in tight gas reservoirs
is a challenge that, if resolved, can lead to an effective strategy to increase recovery. Proppant
hydraulic fracturing is performed to improve well productivity in tight gas reservoirs.
Additionally, a horizontal well can be hydraulically fractured to enhance well productivity in
these reservoirs; however, the results may be disappointing due to a poor understanding of the
performance of fractured horizontal wells in tight gas reservoirs. A series of reservoir simulation
runs were performed given a set of well and reservoir characteristics.
Common factors include reservoir properties, pressure-volume-temperature (PVT), drainage
area, well geometry, and created fracture geometry. The controlling factors include a number
of fractures, reservoir permeability, skin damage and fracture conductivity. The dimensionless
fracture conductivity, FCD, must be carefully determined when designing a fracturing treatment
in tight gas reservoirs.

Absolute Permeability:
Permeability is measured using a pulse-decay technique, allowing permeability measurements
as low as 0.1mD to 10nD. The pulse decay system consists of two gas reservoirs; an upstream
gas reservoir R1, and a downstream gas reservoir R2, differential pressure (P) transducer, and
another transducer to measure downstream pressure.
Since R1 is equal to R2, the decay of the upstream pressure as fluid flows from R1 to R2 is linked
to the increase of downstream pressure at different P as P decreases with time. The flow
rate can be calculated using the known reservoir volumes, the rate of pressure change and fluid
compressibility (Jones, 1997).

Stress-Dependent Measurements:
Relative PermeabilityStress Relationship:
Permeability versus stress analysis provides an indication of the stress sensitivity of fluid flow in
tight gas samples during loading and unloading cycles. Specifically, we examine the dependence
of gas permeability with effective stress. Effective stress is the difference between an external
stress applied to the sample and the internal pore pressure exerted by pore fluids, here we
assume unity of Biots effective stress coefficient.

Effective stress is an important parameter in reservoir depletion and injection. When producing
from a reservoir, pore pressure drop (depressurization) is analogous to increase in horizontal
and vertical stresses applied to the reservoir, representing loading cycle whereas maintaining
pressure in the reservoir by injecting gas, for example, accounts for effective stress decrease,
representing unloading cycle.
The loading cycle is hydrostatic (1=2= 3) where permeability is measured at 2500, 4000,
6000 and 8000 psi. During the unloading cycle, 1 (the vertical stress) is kept constant at 8000
psi and 2= 3 is decreased to 6000, 4000 and 2500 psi. This procedure might be sensitive to
preferential orientations of micro cracks and allowing them to open by the differential stress.
This could result in an observable enhancement of bulk permeability.
Also it has been pointed out that in situ reservoir rocks cannot be under hydrostatic stress and
accordingly, triaxial conditions are more applicable to reservoir conditions e.g. (Dautriat et al.,
2007; Gray et al., 1963; Rhett and Teufel, 1992). A second loading cycle is performed under
hydrostatic conditions to check if permeability recovers back to its original loading behavior.

Ultrasonic Velocity Measurements:


After the permeability measurements, velocity tests were carried out on the same Mam Tor
samples mentioned previously. The core sample is placed in a triaxial core holder in which three
principal stresses (1, 2, 3) are applied. The vertical load accounts for 1 where the force
applied by the load is divided by the area of the core sample. The horizontal stresses, 2 and
3, are maintained using the oil pressure pump. The ultrasonic waves travel from the pulse
generator to the transducers, where the waveforms are displayed and recorded on a digital
oscilloscope.

The travel time through the rock sample can be determined by measuring the difference
between t1 and t0, travel time with and without the sample, respectively. Velocity (V) is
determined using the following relation, where L is the length of the sample, V= L/(t1-t0). The
error in measuring the length of the sample is estimated to be 0.1mm, in addition to an
estimated error of 0.1s when reading the time from the digital oscilloscope. Overall, the
error in velocity is estimated to be about 50 m/s.

Permeability-Stress Relationship
The results reveal that applying differential stress while unloading does not enhance bulk
permeability of Mam Tor siltstone samples. Instead, all unloading cycles are below the loading
curves, figures, which might suggest that microcracks could not be open by differential stress.
Low-porosity (tight) samples are more stress sensitive than high-porosity ones shown in
previous studies e.g. (Al-Harthy 1999 and Holt 1990). Permeability reduction is greater at low
stresses (approximately less than 4000 psi) and the slope decreases at higher stresses whereby
microcracks close. Partially saturated samples are more stress sensitive than dry samples, refer
to figure 4, and have lower gas effective permeability which might be due fluid redistribution
blocking the path of gas, in addition to increasing trapping effect with increasing saturation,
Lake (2005). When evaluating effective stress, it is assumed that Biot-Willis coefficient equals
one. However, from rock and mineral bulk moduli, Biot-Willis coefficient is found to be less
than one and so, this reduces the effect of pore pressure in the overall effective stress.

Velocity-Stress Relationship
Velocity is higher during the unloading cycle as the grains have become more packed
throughout the loading cycle. Dynamic properties such as bulk modulus, shear modulus,
compressibility coefficient,
Vp/Vs, Poissons ratio and Youngs modulus were evaluated. These parameters can help
differentiate between dry and partially saturated samples. The size of the fluid patches should
be much greater than the diffusion length, also known as the wavelength of Biots slow wave
(Lebedev et al., 2009), to enhance P-wave velocity.

Then from the overall discussion the following conclusions were obtained:
If pressure-dependence of permeability k(p) is ignored, erroneous values of permeability, skin
factor and OGIP will be calculated from well test analysis of tight gas reservoirs.
In constant gas rate cases, for both radial and linear reservoir geometry, the plot of pseudo
pressure versus time give a straight line with different slopes for cases with small draw down
during infinite acting flow. For large draw down (larger pressure range) curves start to bending
up and straight line no longer exist. Then, calculated k and s are wrong and depend on the case
(qg considered).
In constant bottom hole pressure cases, radial and linear reservoir geometry, the plot of
pseudo pressure versus time give a straight line during infinite acting flow for all range of pwf
considered, small and large draw down. A straight line with different slope is obtained for each
value of gamma. Results imply that is not possible to identify a stress dependent permeability

from a constant bottom hole pressure draw down scenario. Calculated k and s are wrong and
depend on the case (pwf considered).
From results of reservoir infinite acting period we can conclude that permeability reduction
depends on gamma and reservoir draw down, qg or pwf, and that no correlation can be made.
For finite acting period analysis, in all cases considered with constant qg and pwf, the OGIP is
miscalculated. The level of error on calculations depends on the draw down in the reservoir.
The use of normalized pseudo time was necessary to correct for changes in reservoir
properties.
From results of reservoir finite acting period we can say that no correlation can be made for
calculation of OGIP; results depend on reservoir draw down, qg or pwf and gamma.
The great impact of permeability, skin factor and OGIP calculations are traduced in business
decisions and profitability for the oil company. Miscalculation in the permeability and skin
factor can lead to take wrong decisions regarding well stimulation. That means to invest
additional money to make well stimulation jobs when there are not necessary, and it reduces
the well profitability.
In the case of OGIP calculation, in most of the cases it is sub estimated, calculated values are
lower than the correct value. It can be taken as an advantage; if we consider that additional gas
wells and reserves would be incorporated in the exploitation plan.
In the absence of lab data, this project proves that permeability modulus concept is a good
mathematical approximation to define a relationship for permeability and pore pressure in the
tight gas reservoir.

--------------------------------------

References

Horner, D.R. 1951. Pressure Build-Up in Wells. Proc., Third World Petroleum Congress, Leiden,
Sec. II, 503.
Lee, W.J. and Holditch, S.A. 1981. Fracture Evaluation With Pressure Transient Testing in LowPermeability Gas Reservoirs. J Pet Technol 33 (9): 17761792. SPE-9975-PA. Cinco-Ley, H.,
Samaniego-V., F., and A.N., D. 1978. Transient Pressure Behavior for a Well With a FiniteConductivity Vertical Fracture. SPE J. 18 (4): 253264. SPE-6014-PA.

Agarwal, R.G., Carter, R.D., and Pollock, C.B. 1979. Evaluation and Performance
Prediction of Low-Permeability Gas Wells Stimulated by Massive Hydraulic Fracturing.
J Pet Technol 31 (3): 362372. SPE-6838-PA.
Li M, Xue GQ, Luo BH et al (2009) Pseudosteady-state trinomial deliverability equation
and application of low permeability gas reservoir. Xinjiang Pet Geol 30(5):593595
Rezaee R, Evans B, Rasouli V, Bahrami H, Orsini C (2012) Whicher range tight gas
sands study
Bahrami H, Mohemad Nour J, Alwerfaly Kh, Mutton G, Owais SA, Al-Waley A (2013)
Whicher range field development. Curtin University, Perth, Australia
Al-Hussainy, R., Ramey, H.R. Jr., and Crawford, P.B.: The Flow of Real Gases Through
Porous Media, JPT (May 1966) 624-36; Trans. AIME, 237.
Raghavan, R., Scorer, J. D. T., and Miller, F. G.: An Investigation by Numerical
Methods of the Effect of Pressure-Dependent Rock and Fluid Properties on Well Flow
Tests, SPEJ (June 1972) 267-75; Trans, AIME 253.
Vairogs, J., and Rhoades, V.W. :Pressure Transient Tests in Formations Having StressSensitive Permeability, JPT (Aug. 1973) 965-970; Trans., AIME 255.
Wattenbarger, R. A., El-Banbi A. H., Villegas M. E. and Maggard, J. B.: Productivity
Analysis of Linear Flow Into Fractured Tight Gas Wells, Paper SPE 39931 presented at
the 1998 SPE Rocky Mountain Regional/Low-Permeability Reservoirs Symposium and
Exhibition, Denver, Colorado, Apr. 5-8, 1-15.
Arevalo-Villagran, J. A., Wattenbarger, R. A., Samaniego-Verduzco, F., and Pham, T.T.:
Production Analysis of Long-Term Linear Flow in Tight Gas Reservoirs: Case
Histories, Paper SPE 71516 presented at the 2001 Annual Technical Conference and
Exhibition, New Orleans, LA, Sept. 30 - Oct. 3, 1-15.

http://osp.mans.edu.eg/geotechnical/Ch2.htm

http://petrowiki.org/Tight_gas_reservoirs

Das könnte Ihnen auch gefallen