Sunteți pe pagina 1din 17

URTeC: 1928491

Dynamic Characterization of Unconventional Gas Reservoirs. Field


Cases
Arévalo-Villagrán J.A., Castellanos-Páez F., SPE, PEMEX E&P, Martínez-Romero N.,
SPE, CNH, and Pumar-Martínez F., SPE, CBM.
Copyright 2014, Unconventional Resources Technology Conference (URTeC)

This paper was prepared for presentation at the Unconventional Resources Technology Conference held in Denver, Colorado, USA, 25-27 August 2014.

The URTeC Technical Program Committee accepted this presentation on the basis of information contained in an abstract submitted by the author(s). The contents of this paper
have not been reviewed by URTeC and URTeC does not warrant the accuracy, reliability, or timeliness of any information herein. All information is the responsibility of, and, is
subject to corrections by the author(s). Any person or entity that relies on any information obtained from this paper does so at their own risk. The information herein does not
necessarily reflect any position of URTeC. Any reproduction, distribution, or storage of any part of this paper without the written consent of URTeC is prohibited.

Abstract

Due to the complexity of unconventional gas reservoirs (shale, tight, and coalbed methane), various analytical
models have been developed to estimate formation parameters. This paper presents the results of the modification
and application of some analytical production data models. These analytical models were developed by Arévalo and
Bello. Additionally, dynamic characterization results are presented using a new model that considers the effects of
gas adsorbed in the formation and ultimately improve well production analysis and the evaluation of some formation
parameters. For this study, production data were taken from wells located in the Eagle Ford formation in
northeastern Mexico. The data were softened in order to utilize the Arévalo models for homogenous-isotropic and
low permeability heterogeneous-anisotropic formations. To identify the flow models and formation parameters, the
flow diagnostic plot [m(pi) – m(pwf)] /qg vs. t and the specialized plot [(m(pi) – m(pwf)] /qg vs. √t were used.
Afterwards, the Bello model was applied to analyze double porosity formations and horizontal wells with multiple
fracturing stages. Using the flow diagnostic and specialized plots, four flow regimes were identified (early linear in
the fracture system, bilinear in the matrix-fracture system, linear in the matrix, and boundary-dominated) and the
reservoir parameters were calculated for constant flowing bottomhole pressure. Finally, from these models, an
analytical model was developed that takes into account the effects of gas adsorbed in the formation. This analytical
model derives from the Bump model to consider changes in gas compressibility (cg) and the King model to consider
changes in the gas compressibility factor (z). It also considers Langmuir isotherms, which characterize gas
desorption in the reservoir due to pressure drops. With these models, the dynamic characterization of the
unconventional gas reservoirs was improved, allowing for a more accurate estimation of the reservoir parameters
and subsequently the optimal development and exploitation of these fields.

Introduction

The United States isn’t the only place in the world with large resources in shale oil and gas. In June 2013, the U.S.
Energy Information Administration (EIA) estimated technically recoverable international shale gas across 137
formations in 41 countries (excluding the U.S.) at 6,634 tcf, or nearly ten times the 665 tcf it estimated for the
United States. This is 10% higher than the EIA’s earlier estimates of 5,760 tcf in 2011. The international total
amount excludes some countries in the Middle East, which still have significant conventional natural gas reserves
still in place1.

Shale test wells have already been fracture-stimulated in Argentina, Australia, the United Kingdom, Poland, China,
and Mexico. Some important shale hydrocarbon basins have been identified in Mexico such as La Pimienta-La
Casita and Eagle Ford formations in which we estimate combined Technically Recoverable Resources (TRR) of 545
trillion cubic feet (tcf) of gas, which represents 27% of North American shale gas reserves and 7.5% of shale gas
reserves worldwide. Mexico began exploring its shale basins in 2011. These basins, similar to others, have
geological formations ranging from low permeability (less than 0.1 md) to extremely low permeability (nano-
URTeC 1928491 2

darcies). For this reason, to exploit them, it is necessary to drill horizontal wells with multiple fracking stages in
order to improve the fluid transmissibility from the formations to the producer wells.

According to Clarkson2, various methods have been developed to characterize the production behavior of
unconventional gas reservoirs. These methods include type curves for fractured wells, straight line techniques (flow
regimes), numerical simulation models, and empirical methods, all with or without desorption of adsorbed gas. For
the characterization of unconventional shale gas reservoirs and coalbed methane, it is very important to take
adsorbed gas into account because this is one of the main differences between conventional gas reservoirs and
unconventional low permeability reservoirs (tight gas) since the original gas in place represents an additional
amount of free gas that can be produced. As adsorbed gas is often not taken into consideration, frequently, mistakes
are made in the calculation of this original gas in place and reserves.

Most of the completed vertical and horizontal wells in unconventional gas reservoirs containing one or several
fracking stages generally show long-term transient flows (linear and bilinear) as a consequence of the extremely low
permeability in the formations. Additionally, there are large pressure and gas rate drops in short production times.
As such, an accurate characterization is needed to optimize field development and production.

Conceptual model for unconventional gas reservoirs

Organic material content and adsorbed gas are the governing factors that have a major influence on the behavior of
unconventional gas reservoirs with low permeability. This behavior can be represented through conceptual models
taking the following concepts into consideration: 1) storage mechanisms in unconventional gas reservoirs, 2)
transport in unconventional gas reservoirs, and 3) physical gas adsorption and desorption effects in unconventional
reservoirs. In the analysis and dynamic characterization, these models allow us to better understand these kinds of
reservoirs.

1. Storage mechanisms in unconventional gas reservoirs. There are two main types of gas storage in
unconventional gas reservoirs: free gas in the matrix pores and adsorbed gas in the matrix surface 3. A theoretical
triple-porosity model that includes both free gas and adsorbed gas is used to consider all of the gas that is stored in
formations that contain organic material. A triple porosity system is a combination of a double-porosity system,
matrix fractures, and adsorbed gas in which free gas is stored in the double-porosity system. Matrix micro-pores are
considered porosity 1, natural fractures are porosity 2, and gas adsorbed is porosity 3, although in reality, the storage
of the gas adsorbed is not in the pore-fracture space, it is in the surface of the formation particles in the matrix4 (Fig.
1). Porosity 3 is a virtual porosity that consists of the storage capacity of adhered gas in unconventional reservoirs
that exists due to the lack of permeability in unconventional reservoirs. The reason why this virtual porosity does not
exist in conventional reservoirs is because the permeability in these reservoirs allows for fluid flow. As such, the
organic material content might not be adsorbed into the matrix.

2. Transport in unconventional gas reservoirs. In the primary porosity in unconventional formations that are rich
in organic material, there are large surface areas for gas adsorption that allow for the storage of large amounts of
gas. However, the rock pores are extremely small, which causes the system permeability of this primary porosity to
be substantially small, resulting in practically no gas or water flow5.

King’s gas transport model established that a diffusion process is present in the primary porosity. This diffusion
process can be divided into three different mechanisms a) rock matrix diffusion where molecule-molecule
interactions dominate, b) Knudsen diffusion where molecule-surface interactions dominate, and c) surface diffusion
from the adsorbed gas layer.

Depending on the gas and rock properties, the above mechanisms can act individually or together during the
transport process. The mechanism that dominates the transport in the primary matrix porosity obeys the first law of
Fick (concentration gradient). Darcy Flow is not reached due to the fact that the permeability is very low (equation
1).

−𝐷𝐴 𝑧𝑠𝑐 𝑅𝑇𝑠𝑐 𝑑𝐶 (1)


𝑞𝑔 =
𝑝𝑠𝑐 𝑑𝑥
URTeC 1928491 3

where D is the diffusion coefficient in ft2/day, and C is the molar concentration in lb-mol/ft3. In the case of
secondary porosity in the natural fracture system, diffusion can be present in two stages: 1) there is diffusion from
the matrix to the fractures due to the liberation of the adsorbed gas from the surface of the primary porosity as a
consequence of the pressure drop, and 2) free gas is transported by Darcy Flow into the natural fractures toward the
producer well. The fracture system acts in two ways: a) as an injection system for the primary porosity network and
b) as a conduct for well production5 (Fig. 2).

When gas is desorbed due to a drop in pressure, the gas molecules are able to move and diffuse into the porous space
of the surface particles. The diffusion time is negligible because the formation pores are generally very small in a
micro-scale. Once the gas is liberated from both the surface of the organic material and the rock pores, this gas is
free and follows the same transport mechanisms into the matrix pores and the fracture systems, similar to the
original free gas in the formation3.

3. Physical gas adsorption and desorption in unconventional reservoirs. In unconventional gas reservoirs that
present organic material content, a storage mechanism that is different from conventional gas reservoirs (where the
gas is compressed in the formation pores and fractures) is the additional phenomenon of adsorption of the gas
molecules to the organic rock walls. This is known as physical adsorption or physisorption, a process in which the
electronic structure of the atom or molecule is barely perturbed upon adsorption. In other words, the adsorbed
molecules conserve their chemical nature.

The Langmuir model has been the most commonly used model in the petroleum industry to describe the gas
adsorption phenomenon in solids6, which considers that a gas molecule is adsorbed in a single place and doesn’t
affect neighboring molecules. Additionally, these molecules are desorbed randomly. The Langmuir model can be
represented by equation 2:

𝑉𝐿 𝑝 (2)
𝑉𝑎 =
𝑝𝐿 + 𝑝

where Va is the total volume of the adsorbed gas per volume unit that is in equilibrium in the reservoir at pressure p,
VL is the Langmuir volume or maximum adsorbed volume per volume unit in the reservoir at infinite pressure, and
pL is the Langmuir pressure that represents the pressure in which the adsorbed volume Va is equal to half of the
Langmuir volume, VL. A typical Langmuir isotherm is shown in Fig. 3.

Langmuir isotherms describe the maximum gas amount in an unconventional reservoir that can be stored under
certain conditions with organic material content, pressure and temperature. There are different factors that can
decrease the maximum adsorption capacity of reservoir gas such as organic material content, p, and T. In these
cases, saturated parameters are used in which the desorption or saturation pressure is equal to the undersaturated
initial reservoir pressure, which can be graphically represented by an initial gas content that is below the isotherm
(Fig. 4).

Other main parameters to consider during the exploitation of unconventional gas reservoirs include the Langmuir
parameter values since they determine the type of isotherm and in consequence the desorption pressure, the gas
storage volume, and the desorbed gas that can be produced during exploitation (Fig. 5).

Model modification to consider desorbed gas

The main topic of this work consists of the modification of the equations developed by Arévalo7 and Bello8 for
conventional and unconventional reservoirs in order to take into account the desorption process as a modified
function of pseudotime. The development of these models is presented below.

Bumb and McKee9 showed that the desorption phenomenon can be taken into account in the solutions for the dry
gas diffusion equation, using gas pseudopressure and modified total system compressibility (𝑐𝑡∗ ), represented by
equations 3 and 4, respectively.
URTeC 1928491 4

𝑝
(3)
𝑝
𝑚(𝑝) = 2 ∫ 𝑑𝑝
𝜇𝑧
𝑝0

and

𝑐𝑡∗ = 𝑐𝑔 (1 − 𝑆𝑤 ) + 𝑐𝑤 𝑆𝑤 + 𝑐𝑓 + 𝑐𝑑 (4)

where:

𝑝𝑠𝑐 𝑇𝑉𝐿 𝑝𝐿 𝑧̅ 𝜌𝑔𝑠𝑐 𝑉𝐿 𝑝𝐿 (5)


𝑐𝑑 = =
𝑇𝑠𝑐 𝜙𝑝̅ (𝑝̅ + 𝑝𝐿 )2 𝜙𝜌̅𝑔 (𝑝̅ + 𝑝𝐿 )2

In order to eliminate the nonlinearity of the diffusion equation, Geramy et al.10 and Clarkson11 used the Fraim and
Wattenbarger pseudotime function12, which is expressed in terms of average formation pressure, (𝑝̅), for drawdown
pressure tests or production data. These authors studied flow regimes from a production data analysis point of view
through the derivative function for normalized rate. They defined the term “time match function” taking into account
the definition of modified apparent pseudotime (equation 6) that considers gas desorption effects using (𝑐𝑡∗ ).

𝑡 𝑑𝑡 (6)
𝑡𝑎∗ (𝑝̅ ) = 𝜇𝑖 𝑐𝑡∗ 𝑖 ∫0
(𝜇𝑐𝑡∗ )𝑝
̅

These variables consider instantaneous desorption, which is a reasonable assumption for long-term gas production in
some low-permeability, shale, and coalbed methane reservoirs. The modified pseudotime function can be included
in the multi-fractured horizontal well models, in accordance with Table 1. The desorbed gas effect can be considered
in the pseudotime function to resolve the modified diffusion equation for adsorbed gas 9, 11.

Using this modification to the Arévalo7 and Bello8 models that dynamically characterizes the reservoir using the
normalized rate derivative in the diagnostic log-log plots, flow regimes can be identified and reservoir properties can
be calculated in both conventional and unconventional gas reservoirs. These new models take into account gas
desorption in the pseudotime function and are presented in Tables 2 and 3, for vertical and horizontal wells,
respectively.

Field case analysis

Following is a description of the application of the new models modified by Castellanos and Arévalo to field cases
(presented in this paper) in order to characterize dynamically the reservoir and estimate reservoir parameters and gas
volume. Actual well data were used to characterize and simulate the behavior of a producing well in a shale
formation both with and without adsorbed gas.

Analysis of Shale A well. Table 4 presents data from well located in the Eagle Ford shale formation in southern
Texas13, and the map shows the location of areas according to the fluids produced (Fig. 6). Well Shale A is a
horizontal dry gas production well, which was completed with a stimulation treatment consisting of ten stages of
lateral fracturing at 4,000 ft., resulting in only 20 effective transverse fractures. A Stimulated Reservoir Volume
(SRV) of 169 MMft3 was estimated. The thickness of the production zone is 283 ft. Table 5 presents the general
well data13.

Fig. 7 shows the gas rate and the cumulative gas production history, and Fig. 8 illustrates the bottomhole flowing
pressure log. In both cases, the data correspond to 250 days of production. To detect the flow geometries from the
production data, the normalized gas rate is plotted against time and modified apparent pseudotime (𝑡𝑎∗ ) as shown in
Fig. 9 and considering the case of a well with and without gas desorption, respectively. Two slopes can be observed.
The first slope corresponds (m2 =1/4) of a probable bilinear flow for a period of 5-50 days and the second slope
(m3 = 1/2) corresponds to a later linear flow for a period of 60-200 days. Finally, starting at t = 225 days, it can be
seen how the flow is dominated by the boundary effects.
URTeC 1928491 5

Subsequently, specialized graphs of the normalized rate vs. a specified function of time were created for bilinear and
linear flows (Fig. 10 and 11, respectively). To estimate formation gas desorption parameters and the modified
apparent pseudotime values from Tables 6 and 7 were used 14, 15.
Once the present flow regimes are, the formation parameters and the original gas in place can be estimated. Table 8
shows the matrix and fracture permeabilities without and with gas desorption that were estimated using the
equations to characterize horizontal wells from Bello8 (Table 3) using the slopes values of m3 and m4 obtained from
the specialized plots (Fig. 10 and 11), respectively.

To estimate the Original Gas in Place (OGIP), equation 7 can be used, presented by Wattenbarger et al.16, which is
based on the assumption that boundary dominated flow begins when the pressure in the center of the matrix blocks
begins to decline. The Langmuir equation that quantify the adsorbed gas volume in the producer formation at initial
condition is given as

200.6 𝑇 𝑆𝑔𝑖 √𝑡𝑙𝑟 𝑝𝑖 𝑉𝐿 (7)


𝑂𝐺𝐼𝑃 = ∙ + 𝑚𝑟
(𝜇𝑐𝑡 𝐵𝑔 ) 𝑚 3 𝑝𝑖 + 𝑝𝐿
𝑖

Where tlr is the time in which the straight line stops in the specialized graph of the square root of time. These
allowed estimating OGIP of 3.15 Bscf and 4.06 Bscf without and with gas adsorption, respectively. There is
variation between the values estimated from the permeabilities and the original gas in place since, depending on the
well isotherm (Fig. 9) and the bottomhole pressure, the adsorbed gas is being released to the free gas phase, which
causes the gas compressibility in the formation to be modified along with pseudotime function. Likewise, the
original gas in place considering gas adsorption is modified by an additional 25% due to the gas adsorbed in the
formation. With these results the modified models can be used to characterize the formation and calculate OGIP in
unconventional gas reservoirs considering the gas desorption phenomena.

Analysis of Shale B well. In February 2011, Shale well B was drilled and completed with a horizontal geometry in
Eagle Ford’s upper Cretaceous formation, with a vertical depth of 8,300 ft and a horizontal path of 13,356 ft. During
its completion, 17 fractures were made, 2 more than what had originally been planned in the well design. On
average, the fractures are 856 ft in length, 459 ft in height, and an average width of 0.8 inches. During completion, a
fish was left in hole at 11,030 ft, behind which, according to production logs, there is currently no production. As
such, of the 3,970 ft horizontal stimulated, only 1,837 ft are producing via 8 hydraulic fractures, which is less than
what had been anticipated. Fig. 12 illustrates the mechanical state of Shale B well.

The well turned out to be a producer of dry gas. Fig. 13 illustrates the pressure-production history of the well during
the first 250 days, and Table 9 and Fig. 14 present the general well data and drainage volume of the well. From the
well completion data, we obtain a ye = 246 ft and the following:

𝐻𝑜𝑟𝑖𝑧𝑜𝑛𝑡𝑎𝑙 𝑤𝑒𝑙𝑙 𝑙𝑒𝑛𝑔𝑡ℎ 1837 (8)


𝐿= = = 230 𝑓𝑡
𝐸𝑓𝑓𝑒𝑐𝑡𝑖𝑣𝑒 𝑓𝑟𝑎𝑐𝑡𝑢𝑟𝑒𝑠 8

As such, the cross-sectional area to flow is:


Acw = 2xeh = 2 x 1837 x 492 = 1,807,411 ft2 (9)

Therefore, the matrix–natural fracture area between the blocks formed by the hydraulic fractures is:

Acm = 2 x 2yehL = 2 x 2 x 246 x 1837 x 492 x 8 = 3,873,024 ft2 (10)

For the analysis of the flow regimes and geometries, the normalized rate was plotted against apparent pseudotime
(Fig. 15). Subsequently, due to the fact that linear flow with a slope of m = 1/2 was observed, the specialized graph
for normalized rate vs. √𝑡𝑎∗ was made (Fig. 16).

Using the m3 slope of Fig. 16 and equation 3.3 from Table 3, matrix permeability (km) was estimated, assuming that
the fracture porosity is negligible compared with matrix porosity13,
URTeC 1928491 6

(𝜙𝜇𝑐𝑡 )𝑓+𝑚 = (𝜙𝜇𝑐𝑡 )𝑚 (11)

The reservoir is saturated in its desorption pressure; as such, it is considered that the desorption began with reservoir
production. Additionally, rock compressibility (cf) is negligible in comparison with gas compressibility; therefore,
we have,

𝑐𝑡∗ = cg + cd (12)

Applying the adsorbed gas values from Tables 6 and 10, pertaining to Eagle Ford in the U.S. and published by
Mullen14, the gas volume that is desorbed from the formation was estimated. Following are the diagnostic and
specialized graphs (Figs. 15 and 16), as well as the results obtained in the estimation of parameters using the
modified Bello model in order to consider desorbed gas. To estimate the parameters, instantaneous desorption was
taken into account as if all of the released gas would be produced.

The parameters obtained from the characterization are matrix permeability (km) of 3.85 x 10-6 md, and OGIP of 1.70
Bscf. Fig. 17 shows the results of the gas rate match vs. √𝑡𝑎∗ and the variation of the normalized pseudopressure vs.
√𝑡𝑎∗ with actual data.

In Fig. 17 it can be observed that the match of the pressure-production history data with the modified model by
Castellanos–Arévalo is good. With this model, desorption can be considered in any stage of well production since
the desorbed gas produced can vary depending on the type of isotherm and the initial reservoir pressure and the gas
desorption pressure.

Conclusions

In unconventional gas reservoirs with high organic material content, it is important to consider the gas that is
adsorbed in the formation since this can significantly alter the original gas in place and the estimated parameters
such as primary and secondary permeabilities. The following conclusion can be drawn from this study.

1. It was proven that, by applying Langmuir’s isotherm model, it is possible to take into consideration and
predict the behavior of adsorbed and desorbed gas in unconventional formations that contain organic
material. This is significant since, once gas desorption pressure is reached, there is an additional production
mechanism in the reservoir.
2. It was proven that the pseudotime developed for the characterization of conventional gas reservoirs can be
effectively applied to unconventional formations, taking into account instantaneous gas desorption in the
total compressibility of the system and depending on the average reservoir pressure.
3. Through the analysis of the well data, it was possible to confirm the applicability of the modified models to
analyze production and characterization data from unconventional gas reservoirs, taking into consideration
the phenomenon of adsorption using Langmuir’s isotherm and modified pseudotime at any moment during
well production.
4. For the characterization models used in this work, the assumption was made that the desorbed gas is
instantaneous, obtaining good results. However, it is important to bear in mind that desorption is not
instantaneous in all reservoirs. As such, it is recommended to adjust the models taking into account real gas
desorption time.
5. Another important point to keep in mind regarding the adsorption and desorption models is that Langmuir
isotherms only consider the monocomponent fluid, methane gas. However, it has been observed that in
some reservoirs there are multicomponent blends. As such, it is recommendable to utilize the
multicomponent Langmuir isotherm or to study how to adjust a cubic state equation, allowing us to better
characterize the desorption phenomenon.

Acknowledgments

We would like to thank PEMEX E&P for permission to present the data and results of some tight and shale wells.
We are also thankful to Cynthia Sperry for her efficient work and participation in preparing this paper.
URTeC 1928491 7

Nomenclature
A= area, L2, ft2
Acm = matrix-natural fracture area between the blocks formed by the hydraulic fractures, L2, ft2
Acw = cross –sectional area to flow, L2, ft2
bSSPDP = intercept for boundary-dominated flow of ∆𝑚(𝑝)/𝑞𝑔 vs. t plot, psia2-D/Mscf-cp
C= molar concentration, m/L3, lb-mol/ft3
cd = desorbed gas compressibility, Lt2/m,1/psia
cf = formation compressibility, Lt2/m,1/psia
CA = shape factor
cg = gas compressibility, Lt2/m,1/psia
ct = total system compressibility, Lt2/m,1/psia
𝑐 t* = modified total system compressibility, Lt2/m,1/psia
cw = water compressibility, Lt2/m,1/psia
D= diffusion coefficient, L2/t, ft2/day
EIA = Energy Information Administration
k= formation permeability, md
kf = fracture permeability, L2, md
km = matrix permeability, L2, md
L= effective producing length in the horizontal section of the well, L, ft
m= straight line slope
mr = rock mass
m(p) = real gas pseudo-pressure at average reservoir pressure, m/Lt3, psia2/cp
m(pwf) = pseudo-pressure at bottomhole flowing pressure, m/Lt3, psia2/cp
mLPC = slope for constant pwf for bilinear or linear flows from specialized plot, psia2-D/Mscf-cp
mCPRPC = slope for constant pwf for radial flow from specialized plot, psia2-D/Mscf-cp
mCRSDP = slope for constant qg for spherical flow from specialized plot, psia2-D/Mscf-cp
p= pressure, m/Lt2, psia
𝑝̅ = average reservoir pressure, m/Lt2, psia
pL = The Langmuir pressure that represents the pressure in which the adsorbed volume V a is equal
to half of the Langmuir volume, VL, m/Lt2, psia
PSS = pseudostationary state flow
qg = gas production rate, L3/t, Mscf/D
R= gases universal constant, equal to 10.732 (psia -ft3)/ (lbm-mol-°R)
rw = well radius, L, inches
S= skin factor, adim
SRV = Stimulated Reservoir Volumen, L3, MMft3
Sw = water saturation, %
t= time, days
T= reservoir temperature, °R
𝑡𝑎∗ = modified apparent pseudotime, t, days
tlr = time in which the straight line stops in the specialized graph of the square root of time, t, days
TRR = Technically Recoverable Resources, L3, tcf
TVD = true vertical depth, L, ft
Va = total volume of the adsorbed gas per volume unit that is in equilibrium in the reservoir at
pressure p, L3, ft3
VL = The Langmuir volume or maximum adsorbed volume per volume unit in the reservoir at
infinite pressure, L3, ft3
w= width of the fracture, L, inches
x= distance, ft
ye = distance from the well to the formation boundary, L, ft
z= real gas compressibility factor
ρ= density, m/L3, lb/ft3
µ= viscosity, m/Lt, cp
φ= porosity, %
ω= acentric factor
URTeC 1928491 8

tcf = trillion cubic feet

Subscripts
𝑎= apparent
cw = cross sectional area to flow
cm = matrix-natural fracture area between the blocks formed by the hydraulic fractures
d= desorbed
f= fracture
m= matrix
mf = matrix-fracture system
sc = standard conditions
g= gas
i= initial
r= rock
t= total
w= water

References

1. Gruber, S., Beveridge, N., Brackett, B., Clint, O., Rawat, H. and Shiah, H. (Dec. 10, 2013). Bernstein
Energy: International Shale – Hope Shifts but Inherent Challenges Continue to Stymie Development,
Bernestein Research.
2. Clarkson, C. R., Jensen, J. L., & Blasingame, T. (2011, January 1). Reservoir Engineering for
Unconventional Reservoirs: What Do We Have to Consider? Society of Petroleum Engineers.
doi:10.2118/145080-MS.
3. Lane, H. S., Lancaster, D. E., & Watson, A. T. (1990, January 1). Estimating Gas Desorption Parameters
From Devonian Shale Well Test Data. Society of Petroleum Engineers. doi:10.2118/21272-MS.
4. Bo Song (2010). Pressure transient analysis and production analysis for New Albany shale gas wells,
Master of Science thesis, Texas A&M University.
5. King, G. R. (1990, January 1). Material Balance Techniques for Coal Seam and Devonian Shale Gas
Reservoirs. Society of Petroleum Engineers. doi:10.2118/20730-MS.
6. Langmuir, I. (1918), The Adsorption of Gases on Plane Surfaces of Glass, Mica and Platinum, Am. Chem.
Soc. 40, 1361.
7. Arevalo Villagran J. (2001). Analysis of long-term behavior in tight gas reservoirs: Case histories. PhD
Dissertation, Texas A &M U, College Station, Texas, E.U.
8. Bello, R.O. (2009, May). Rate Transient Analysis in Shale Gas Reservoirs with Transient Linear
Behaviour. PhD Dissertation, Texas A &M U, College Station, Texas, E.U.
9. Bumb, A.C., & McKee, C.R. (1988). Gas-Well Testing in the Presence of Desorption for Coalbed Methane
and Devonian Shale. SPEFE 3 (1): 179 – 185. SPE-15227-PA. doi: 10.2118/15227-PA.
10. Gerami, S., Pooladi-Darvish, M., Morad, K., & Mattar, L. (2008, July 1). Type Curves for Dry CBM
Reservoirs With Equilibrium Desorption. Petroleum Society of Canada. doi:10.2118/08-07-48.
11. Clarkson C.R.(2012). “Rate Transient Analysis of 2-Phase (Gas + Water) CBM Wells”, Journal of Natural
Gas Science and Engineering 8, pg. 106-120.
12. Fraim, M. L., & Wattenbarger, R.A. (1987, December 1). Gas Reservoir Decline-Curve Analysis Using
Type Curves With Real Gas Pseudopressure and Normalized Time. Society of Petroleum Engineers.
doi:10.2118/14238-PA.
13. Xu, B., Haghighi, M., Cooke, D. A., & Li, X. (2012, January 1). Production Data Analysis in Eagle Ford
Shale Gas Reservoir. Society of Petroleum Engineers. doi:10.2118/153072-MS.
14. Mullen, J. (2010, January 1). Petrophysical Characterization of the Eagle Ford Shale in South Texas.
Society of Petroleum Engineers. doi:10.2118/138145-MS
15. Andrews, I.J. (2013). The Carboniferous Bowland Shale Gas Study: Geology and Resource Estimation.
London, UK, British Geological Survey for Department of Energy and Climate Change, Apendix A, pg. 03.
16. Wattenbarger, R. A., El-Banbi, A. H., Villegas, M. E., & Maggard, J. B. (1998, January 1). Production
Analysis of Linear Flow Into Fractured Tight Gas Wells. Society of Petroleum Engineers.
doi:10.2118/39931-MS.
URTeC 1928491 9

Table 1. Flow regimes identification through the derivative function for normalized rate and pseudotime function 2, 4.

Flow regime Log-Log diagnostic Derivative


function Type of plot
slope
Bilinear flow |𝑚(𝑝𝑖 ) − 𝑚(𝑝𝑤𝑓 )| (1.1)
Radial derivative m = 1/4 𝑣𝑠. 𝑙𝑜𝑔(𝑡 ∗ )
𝑞𝑔

|𝑚(𝑝𝑖 ) − 𝑚(𝑝𝑤𝑓 )| (1.2)


Bilinear derivative m=0 4
𝑣𝑠. √𝑡 ∗
𝑞𝑔
Linear flow |𝑚(𝑝𝑖 ) − 𝑚(𝑝𝑤𝑓 )| (1.3)
Radial derivative m = 1/2 𝑣𝑠. 𝑙𝑜𝑔(𝑡 ∗ )
𝑞𝑔

Linear derivative m=0 |𝑚(𝑝𝑖 ) − 𝑚(𝑝𝑤𝑓 )| (1.4)


𝑣𝑠. √𝑡 ∗
𝑞𝑔
Radial flow (1.5)
|𝑚(𝑝𝑖 ) − 𝑚(𝑝𝑤𝑓 )|
Radial derivative m=0 𝑣𝑠. 𝑙𝑜𝑔(𝑡 ∗ )
𝑞𝑔

|𝑚(𝑝𝑖 ) − 𝑚(𝑝𝑤𝑓 )| (1.6)


PSS flow.
Radial derivative m=1 𝑣𝑠. 𝑙𝑜𝑔(𝑡 ∗ )
𝑞𝑔

Table 2. New models of the derivative function for normalized gas rate for vertical wells 7 (modified by Castellanos
and Arévalo, 2014) to take into account gas desorption in the pseudotime function.

Flow regime Specialized plot Interpretation equation


𝑛
(𝑞𝑔𝑗 − 𝑞𝑞𝑗−1 ) ∗ |𝑚(𝑝𝑖 ) − 𝑚(𝑝𝑤𝑓 )| 𝛼𝑇 1 (2.1)

Lineara ∑ √𝑡𝑎𝑛 − 𝑡𝑎(𝑛−1) 𝑣𝑠. √𝑘𝑚𝑓 𝐴𝑐 = ( )
𝑗=𝑖
𝑞𝑔𝑛 𝑞𝑔 √𝜇𝑔𝑖 (𝜙𝑉𝑐𝑡 )𝑓𝑖 + (𝜙𝑉𝑐𝑡 )𝑚𝑖 𝑚𝐿𝑃𝐶

𝑛
(𝑞𝑔𝑗 − 𝑞𝑞(𝑗−1) ) 4 ∗ |𝑚(𝑝𝑖 ) − 𝑚(𝑝𝑤𝑓 )| 3⁄ 984 4√𝐴𝑐 𝑇 1 (2.2)
Bilinear b ∑ ∗
√𝑡𝑛 − 𝑡𝑛−1 𝑣𝑠. 𝑘𝑚𝑓4 √𝑤 = ( )
𝑞𝑔𝑛 𝑞𝑔 4 𝑚𝐿𝑃𝐶
𝑗=𝑖 √𝜇𝑔𝑖 [(𝜙𝑉𝑐𝑡 )𝑓𝑖 + (𝜙𝑉𝑐𝑡 )𝑚𝑖 ]
𝑛
(𝑞𝑔𝑗 − 𝑞𝑞(𝑗−1) ) |𝑚(𝑝𝑖 ) − 𝑚(𝑝𝑤𝑓 )| 1640𝑇 (2.3)
∗ ∗
Radialc ∑ 𝑙𝑜𝑔(𝑡𝑎𝑛 − 𝑡𝑎(𝑛−1) ) 𝑣𝑠. (𝑘ℎ)𝑚𝑓 =
𝑞𝑔𝑛 𝑞𝑔 𝑚𝐶𝑃𝑅𝑃𝐶
𝑗=𝑖
𝑛 2
(𝑞𝑔𝑗 − 𝑞𝑞(𝑗−1) ) 1 |𝑚(𝑝𝑖 ) − 𝑚(𝑝𝑤𝑓 )| (2.4)
Spherical b ∑ 𝑣𝑠. 10098 3
𝑞𝑔𝑛 𝑞𝑔 𝑘𝑓 = [√𝜇𝑔𝑖 [(𝜙𝑉𝑐𝑡 )𝑓𝑖 + (𝜙𝑉𝑐𝑡 )𝑚𝑖 ] ( )]
𝑗=𝑖 ∗ − 𝑡∗
√𝑡𝑎𝑛 𝑎(𝑛−1)
𝑚𝐶𝑅𝑆𝐷𝑃

Boundary 𝑛
(2.5)
(𝑞𝑔𝑗 − 𝑞𝑞(𝑗−1) ) ∗ ∗
|𝑚(𝑝𝑖 ) − 𝑚(𝑝𝑤𝑓 )| 712𝑇 2.2458 𝐴
dominated ∑ (𝑡𝑎𝑛 − 𝑡𝑎(𝑛−1) ) 𝑣𝑠. (𝑘ℎ)𝑓 = [𝑙𝑛 ( ) + 2𝑆]
𝑞𝑔𝑛 𝑞𝑔 𝑏𝑆𝑆𝑃𝐷𝑃 𝐶𝐴 𝑟𝑤2
effectsb 𝑗=𝑖

Solution:
a. 𝐶𝑜𝑛𝑠𝑡𝑎𝑛𝑡 𝑝𝑤𝑓 , 𝛼 = 201 and constant 𝑞𝑔 , 𝛼 = 128
b. For constant 𝑞𝑔
c. For 𝑝𝑤𝑓 and constant 𝑞𝑔
URTeC 1928491 10

Table 3. New models of the derivative function for normalized gas rate for horizontal wells 8 (modified by
Castellanos and Arévalo, 2014) to take into account gas desorption in the pseudotime function.

Derivative
Flow regime function Type of plot Interpretation equation
plot
|𝑚(𝑝𝑖 ) − 𝑚(𝑝𝑤𝑓 )| 1262𝑇 1
Early linear flow m = 1/2 𝑣𝑠. √𝑡𝑎∗ √𝑘𝑓 𝐴𝑐𝑤 = (3.1)
𝑞𝑔 𝑚
√𝜔(𝜙𝜇𝑐𝑡 )𝑓+𝑚 1

|𝑚(𝑝𝑖 ) − 𝑚(𝑝𝑤𝑓 )| 4064𝑇 1


Bilinear flow m = 1/4 𝑣𝑠. 4√𝑡𝑎∗ √𝑘𝑓 𝐴𝑐𝑤 = 0.25 𝑚 (3.2)
𝑞𝑔 [𝜎𝑘𝑚 (𝜙𝜇𝑐𝑡 )𝑓+𝑚 ] 2

Linear flow |𝑚(𝑝𝑖 ) − 𝑚(𝑝𝑤𝑓 )| 1262𝑇 1


m = 1/2 𝑣𝑠. √𝑡𝑎∗ √𝑘𝐴𝑐𝑤 = (3.3)
𝑞𝑔 √(𝜙𝜇𝑐𝑡 )𝑓+𝑚 𝑚3

Transitory linear
|𝑚(𝑝𝑖 ) − 𝑚(𝑝𝑤𝑓 )| 1262𝑇 1
flow in the m = 1/2 𝑣𝑠. √𝑡𝑎∗ √𝑘𝑚 𝐴𝑐𝑤 = (3.4)
𝑞𝑔 √(𝜙𝜇𝑐𝑡 )𝑚 𝑚4
matrix

Table 4. Data from a well located in the Eagle Ford shale formation in southern Texas 13.

Depth, ft 2,500 - 14,000

Net thickness, ft 50 - 300

Pressure gradient, psia/ft 0.4 - 0.8

TOC, % 2-9

Gas saturation, %: 83 – 85

Permeability, nd 1 - 800

Table 5. Data from Shale A well13.

Well radius, ft 0.33


Lateral length, ft 4,000
Thickness, ft 283
Depth, TVD, ft 10875
Hydrocarbon porosity (%) (φhc = φef (1-Sw)) 5.76
Reservoir pressure, psia 8,350
Temperature, °R 745
Gas compressibility, 10-5 psia-1 6
Gas viscosity, cp 0.03334
Number of effective fractures 20
Stimulated Reservoir Volume (SRV), MMft3 169
URTeC 1928491 11

Table 6. Gas adsorbed values in shale formation in the U.S.15.

Tabla 7. Rock parameters to estimate gas desorption in Shale A well 13.

VL = 720 scf/Ton ρr = 2.5 gr/cm3


PL = 550 SRV = 16900000 ft
T= 285 °F mr = 1197306.06 Ton
φ= 0.0576

Table 8. Matrix and fracture permeabilities for Shale A well for both without and with gas desorption.

Permeability Without gas desorption With gas desorption


matrix (𝑘𝑚 ): 2.15 × 10−4 𝑚𝑑 1.28 × 10−5 𝑚𝑑
fracture (𝑘𝑓 ): 1.61 × 10−2 𝑚𝑑 2.62 × 10−2 𝑚𝑑

Table 9. General data from Shale B well.

Well radius, ft 0.375


Lateral length, ft 1837
Thickness, ft 492
Depth, TVD, ft 2530
Hydrocarbon porosity (%) (φhc = φef (1-Sw)) 6.0
Reservoir pressure, psia 5,100
Temperature, °R 667
Gas compressibility, 10-4 psia-1 1.3
Gas viscosity, cp 0.0239
Number of effective fractures 8
Stimulated Reservoir Volume (SRV) (MMft3) 445

Table 10. Rock parameters to estimate gas desorption en Eagle Ford 14, 15.
VL = 60 scf/ton ρr = 2.8 gr/cm3
PL = 250 SRV = 446 MMft3
T= 207 °F mr = 35280000 Ton
φ= 0.06
URTeC 1928491 12

Fig. 1. Triple porosity storage model for unconventional gas reservoirs 4.

Fig. 2. Transport sketch in unconventional gas reservoirs with adsorption process 5.

Fig. 3. Langmuir typical isotherm6.


URTeC 1928491 13

Fig. 4. Langmuir isotherms in a gas reservoir: a) saturated in adsorbed gas (left sketch), and b) undersaturated in
adsorbed gas (right sketch).

Fig. 5. Behavior of the adsorption isotherm upon changing the Langmuir pressure.

Fig. 6. Extension of the Eagle Ford formation in southern Texas 13.


URTeC 1928491 14

Fig. 7. Gas rate and cumulative production from well A13.

Fig. 8. Bottomhole flowing pressure history for well A13.

|𝑚(𝑝𝑖 )−𝑚(𝑝𝑤𝑓 )| |𝑚(𝑝𝑖 )−𝑚(𝑝𝑤𝑓 )|


Fig. 9. Diagnostic plots of normalized pseudopressure: a) vs. t (left sketch) and b) vs.
𝑞𝑔 𝑞𝑔
𝑡𝑎∗ (right sketch) detecting both bilinear and linear flows.
URTeC 1928491 15

Fig. 10. Specialized plots to characterize bilinear flow:


|𝑚(𝑝𝑖 )−𝑚(𝑝𝑤𝑓 )| |𝑚(𝑝𝑖 )−𝑚(𝑝𝑤𝑓 )| 0.25
a) vs. 𝑡 0.25 (left sketch) and b) vs. 𝑡𝑎∗ (right sketch).
𝑞𝑔 𝑞𝑔

Fig. 11. Specialized plots to characterize linear flow:


|𝑚(𝑝𝑖 )−𝑚(𝑝𝑤𝑓 )| |𝑚(𝑝𝑖 )−𝑚(𝑝𝑤𝑓 )| 0.50
a) vs. 𝑡 0.50 (left sketch) and b) vs. 𝑡𝑎∗ (right sketch)
𝑞𝑔 𝑞𝑔

Fig. 12. Mechanical state of Shale B well.


URTeC 1928491 16

Fig. 13. Pressure-production history of Shale B well.

Fig. 14. Drainage volume of Shale B well.

|𝑚(𝑝𝑖 )−𝑚(𝑝𝑤𝑓 )|
Fig. 15. Diagnostic plot of vs. 𝑡𝑎∗ for Shale B well.
𝑞𝑔
URTeC 1928491 17

|𝑚(𝑝𝑖 )−𝑚(𝑝𝑤𝑓 )|
Fig. 16. Specialized plot of vs. √𝑡𝑎∗ for Shale B well.
𝑞𝑔

|𝑚(𝑝𝑖 )−𝑚(𝑝𝑤𝑓 )|
Fig. 17. History matches: a) gas rate (qg) vs. √𝑡𝑎∗ y b) vs. √𝑡𝑎∗ .
𝑞𝑔

S-ar putea să vă placă și