Papers Spe 166173 Ms

Embed Size (px)

Citation preview

  • 7/25/2019 Papers Spe 166173 Ms

    1/17

    SPE 166173

    A New Approach for the Simulation of Fluid Flow in UnconventionalReservoirs through Multiple Permeability ModelingBicheng Yan, SPE, Masoud Alfi, SPE, Yuhe Wang, SPE, John E. Killough, SPE, Texas A&M University

    Copyright 2013, Society of Petroleum Engineers

    This paper was prepared for presentation at the SPE Annual Technical Conference and Exhibition held in New Orleans, Louisiana, USA, 30 September2 October 2013.

    This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents of the paper have not beenreviewed 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, itsofficers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written consent of the Society of Petroleum Engineers is prohi bited. Permission toreproduce 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.

    AbstractShale reservoirs are characterized by ultra-low permeability, multiple porosity types, and complex fluid storage and flow

    mechanisms. Consequentially the feasibility of performing simulations using conventional Dual Porosity Models based on

    Darcy flow has been frequently challenged. Additionally, tracking of water in shale continues to be controversial and

    mysterious. In organic-rich shale, kerogen is generally dispersed in the inorganic matter. Kerogen is different from any other

    shale constituents because it tends to be hydrocarbon-wet, abundant in nanopores, fairly porous and capable of adsorbing gas.

    However, the inorganic matter is usually water wet with low porosity such that capillary pressure becomes the dominant

    driving mechanism for water flow, especially during hydraulic fracturing operations. This work presents a technique of

    subdividing shale matrices and capturing different mechanisms including Darcy flow, gas diffusion and desorption, and

    capillary pressure. The extension of this technique forms a solid and comprehensive basis for a specially-designed simulator

    for fractured shale reservoirs at the micro-scale.

    Through the use of this unique simulator, this paper presents a micro-scale two-phase flow model which covers three

    continua (organic matter, inorganic matter and natural fractures) and considers the complex dynamics in shale. In the model,

    TOC is an indispensable parameter to characterize the kerogen in the shale. A unique tool for general multiple porositysystems is used so that several porosity systems can be tied to each other through arbitrary connections. The new model

    allows us to better understand the complex flow mechanisms and to observe the water transfer behavior between shale

    matrices and fractures under a microscopic view. Sensitivity analysis studies on the contributions of different flow

    mechanisms, kerogen properties, water saturation and capillary pressure are also presented.

    Introduction

    In recent years, unconventional resources have played a significant part to balance between the increasing energy demand

    and the shortage of production from the conventional reservoirs in the United States (Wei et al. 2013). Hydrocarbon from

    organic rich shale is one of the most significant unconventional resources. The successful development of shale reservoirs is

    greatly attributed to horizontal well drilling and hydraulic fracturing operations. In industry, effective hydraulic fracturing for

    shale wells is performed mainly through injecting slickwater under high pressure. However, generally the recovery of

    fracturing fluid is quite low.King (2012)suggested that the water might be trapped in the small pores and the micro-fractures

    of shale. Besides, evidence shows that there is a high concentration of chloride salts in the flowback fluid, while it cannot beexplained either from the composition of fracturing fluid (mostly fresh water) or from the constituents of shale and the

    salinity of formation brine (King 2012).Wang and Reed (2009)propose that there exist four pore systems in the organic-rich

    shale: inorganic matter, organic matter (kerogen), natural fractures and hydraulic fractures. It is also suggested that the

    organic matter is oil wet and that single oil or gas phase flow without residual water is dominated in kerogen fragments.

    However, the inorganic matrix is mostly considered as water-wet (Kalakkadu et al. 2013). Through the approach of

    Molecular Dynamics Simulation and with the initial condition of water and NaCl,Hu et al. (2013)suggested that water could

    be filled in the larger pores in the kerogen through capillary condensation but no water enters the smaller 0.9 nm kerogen

    pores. Water exists in the inorganic MgO pores in the liquid phase; meanwhile, there is a much higher ionic concentration in

    the inorganic matter than that in the kerogen.

    In shale gas reservoirs, the source of shale gas can be thermogenic, biogenic or combined source (Darishchev et al. 2013).

    Natural gas is usually considered to exist in three forms: compressed gas in pores and fissures, adsorbed gas in the organic

    and inorganic matter, and dissolved gas in the kerogen (Javadpour 2009;Zhang et al. 2012). Usually it might be reasonable

  • 7/25/2019 Papers Spe 166173 Ms

    2/17

    2 SPE 166173

    that the sorbed gas in the inorganics be neglected, because under moisture conditions methane molecules are preferentially

    sorbed on the walls of hydrophobic pores in the kerogen. In the inorganic matter, water molecules might either block the

    hydrophilic pores or occupy those sorption sites (Ji et al. 2012; Zhang et al. 2012). Using gas solution and diffusion

    parameters from bitumen,Swami and Settari (2012)considered gas solution in the kerogen through an ideal model with a

    cylindrical nanotube surrounded by the kerogen solid bulk and observed the difference caused by solution gas. However, it

    might not be pragmatic to isolate nanopores in the kerogen from the kerogen solid bulk in a conventional simulation based on

    porous media, unless it can be upscaled properly.

    Because the permeability in the organic rich shale is extremely low and the pore size in the shale is in the nanometerscale, some non-Darcian mechanisms including gas diffusion, desorption and slippage flow have been brought into

    consideration in order to explain the profitable gas rate in shale reservoirs (Civan et al. 2011; Ertekin et al. 1986; Freeman et

    al. 2012; Javadpour 2009; Javadpour et al. 2007; Shabro et al. 2011; Shabro et al. 2012). A common approach for those

    authors is to use an apparent permeability either based on mechanisms or through a condition function of Knudsen number.

    Matrix subdivision due to the difference between the kerogen and the inorganic matter is generally not taken into account,

    and the complexity of the connectivities between different pore systems cannot be properly represented. Therefore, it might

    be difficult to observe the dynamics of fluid flow from the shale matrix to outer fractures through those models.

    Previously we have established a single gas phase Micro-Scale Model (Yan 2013; Yan et al. 2013b; Yan et al. 2013a). It

    subdivided the shale matrix into the inorganic matrix and the organic matrix with different pore geometries. The issue of

    connectivities between different continua is solved through the random distribution of kerogen units. The model also

    considers adsorbed gas and gas diffusion in the kerogen and a petrophysical parameter - the content of total organic carbon

    (TOC, wt%). It demonstrates that desorption can maintain a high gas in place level and diffusive flow is quite significant to

    gas rate with permeability down to nanoDarcy magnitude.

    This work is based on our previous research, with the same structure in the micro model (Yan 2013; Yan et al. 2013a), but

    gas and water two phase flow with mixed wettability in the shale matrix is implemented. The effect of high capillary pressure

    is also taken into account here. The work is motivated to interpret the dynamics of gas and water flow at the micro-scale

    level. Case analysis is introduced to evaluate the influence of different mechanisms or parameters on the gas and water flow

    in the model.

    Model Description

    To better catch the dynamics of complex processes during hydrocarbon production from unconventional reservoirs, we have

    used a multiple porosity model previously offered by the authors (Yan et al. 2013a). Supported by different petrophysical and

    geological data, this new model allows us to understand the complex mechanisms and eventually to better predict ultimate

    recovery from shale reservoirs.

    In this paper, four different pore systems are considered with distinctive characteristics. To account for the presence of

    induced or natural fractures in the shale, we have divided the media into fracture (representing natural fractures in thereservoir) and bulk matrix blocks. Characterized by their high permeability, natural fractures serve as pathways to connect

    low permeability shale matrix blocks with the induced fracture network or the well-bore. Based on the difference between the

    kerogen and the inorganic matter in the shale matrix, the matrix bulk is further divided into two sub-media: the organic and

    the inorganic matter (Wang and Reed 2009). As reflected in the geological data (Fig. 1andFig. 2), the presence of different

    pore sizes has inspired us to further subdivide the organic matter into kerogen with micropores and kerogen with nanopores.

    This further subdivision will help us to better understand the effect of different transportation mechanisms in each sub-

    continuum.

    Fig. 1Different pore geometry in shale, especially in kerogen, large and small pores can be observed (InGrain 2010).

  • 7/25/2019 Papers Spe 166173 Ms

    3/17

    SPE 166173 3

    Fig. 2Local pore distribution in kerogen, smaller pores reside on the wall of larger pores (Curtis et al. 2010).

    Considering that shale matrices are surrounded by fractures,Yan et al. (2013a)proposed a configuration of different pore

    systems in a Micro-Scale Model. Specifically each block of kerogen with nanopores is surrounded by six blocks of kerogen

    with micropores in an attempt to mimic the effect that kerogen with micropores ties kerogen with nanopores to any other

    non-kerogen pore systems. This type of grid configuration is supported by the geological and petrophysical data from shale

    rocks, which indicates that in the kerogen those nanopores are located through the wall of the larger pores (Fig. 2). Kerogen

    units are randomly distributed in the matrix bulk through a rigorous Monte Carlo algorithm, and their abundance depends on

    different factors such as TOC and different media properties (Fig. 3). Based on the grid configuration, kerogen with

    nanopores can only connect to kerogen with micropores. In the larger pore sizes, fluid in kerogen with micropores, inorganic

    matter, and natural fractures can either flow in themselves or among each other. Therefore, there should be seven types of

    connections for those pore systems mentioned above (Fig. 4).

    Fig. 3A representative micro-model with random kerogen distribution (green grids: kerogen, other empty space in the cuboid:inorganic matter)(Yan et al. 2013b)

    Fig. 4Schematic of connectivities in the Micro-Scale Model

    Advective flow

    Diffusive flow

  • 7/25/2019 Papers Spe 166173 Ms

    4/17

    4 SPE 166173

    In the Micro-Scale Model, the previous single gas phase flow (Yan et al. 2013a)has been extended to two phase gas

    water flow according to the characteristics of each pore system. As the mean free path of gas molecules becomes somehow

    comparable to the pore diameter, in addition to viscous flow (Darcys law), molecule/walls interactions will play a more

    significant effect so that diffusive flow cannot be neglected. According to Ficks law, the mass flux is p roportional to the

    concentration gradient. Considering the nanometer pore size in kerogen, diffusive flow is considered there. Since the mean

    pore diameter in the kerogen with nanopores is very small, Darcy flow is not considerable there. Therefore, in kerogen with

    nanopores, diffusive flow is assumed to be the only flow mechanism. On the other hand, Darcy flow exists in kerogen with

    micropores, the inorganic matter and the fracture network. Diffusive flow is considered to only occur in the gas phase; waterflow takes place only through Darcy flow.

    Unlike conventional gas reservoirs, the gas storage mechanism in shale reservoirs is not limited to compressed gas;

    adsorbed gas is also important. As we previously mentioned, in a gas-water two phase system, considering the different

    wettability of the inorganic matter and the kerogen, gas adsorption in the kerogen should be considered, however, gas

    adsorption in the inorganic matter might be so weak as to be neglected under moisture conditions. As reservoir pressure

    decreases, adsorbed gas in the kerogen will be gradually desorbed as free gas and flow into fractures. Therefore, the

    desorption process actually increases the gas accumulation within the kerogen grids in this model. A popular form of

    desorption modeling is the Langmuir isotherm (Cui et al. 2009), which has been used in this paper as well.

    In addition to fluid flow and storage mechanisms, in two phase water and gas flow, wettability, relative permeability and

    capillary pressure further increase the complexity of the model. Particularly, capillary pressure might play a significant effect

    due to the small pore size in the shale matrix. Pore systems in the kerogen (micro- and nano-pores) are considered to be gas

    wet (Passey et al. 2010;Wang and Reed 2009)and gas can flow within the kerogen in a continuous phase. However, in the

    inorganic matter, especially the siliclastic minerals are water-wet so that water can be layered on the grain surfaces (Passey et

    al. 2010). Water dynamics in the shale reservoir are important during hydraulic fracturing and the early gas production

    period. Because of its affinity to the inorganic matter and the high capillary pressure between the gas and water phases, water

    is likely to be imbibed into the shale matrix. Although a large quantity of research has studied different aspects of capillary

    pressure in shale reservoirs, this phenomenon still stays ambiguous due to the complex interactions of water molecules with

    extremely small scale pores in the tight gas and shale gas reservoirs. These ultra-small pore sizes, however, can be interpreted

    as a medium with high capillary pressure (Wang et al. 2013). In this paper, the Brooks and Corey formulation (Brooks and

    Corey 1964)is used to calculate capillary pressure.

    c

    ee

    P

    PS (1)

    Wherewr

    wrw

    e S

    SS

    S

    1

    Fig. 5 shows the capillary pressure curve used for the base case simulation in this paper. Due to extremely low absolute

    permeability of shale, it is difficult to physically measure the actual relative permeability in this type of reservoirs. For our

    simulation purposes, Eq. (2) and (3) are used to build non-wetting phase relative permeability curves for the drainage and

    imbibition process. For the wetting phase, however, we have assumed no hysteresis occurs, which means that we can use the

    same formula (Eq. (4)) for both drainage and imbibition cases (Dacy 2010; Molina 1980). Fig. 6 shows the imbibition

    relative permeability curve we have incorporated into our simulator, and Table 1 represents the values used to calculate

    hydraulic properties of the base case or default values.

    nwdn

    wcnwnwc

    wcnwwrnwd

    SS

    SSk

    1

    1 (2)

    nwin

    nwtwi

    wiwrnwi

    SS

    SSk

    11 (3)

    wn

    wcw

    wcwwrw

    S

    SSk

    1 (4)

    Considering that water molecules are just present in the aqueous phase and hydrocarbon molecules can be found only in

    the gaseous phase (solubility of hydrocarbon molecules in the aqueous phase is considered to be negligible in this work), the

    gas phase mass balance equation can be used for the hydrocarbon component and the aqueous phase mass balance can be

  • 7/25/2019 Papers Spe 166173 Ms

    5/17

    SPE 166173 5

    used for the water component. Therefore, the mass balance equations for the two phase micro model can be written as Eq. (5)

    and (6).

    Gas phase:

    t

    q

    t

    SzgP

    KKPDC a

    gg

    gg

    g

    rg

    gg

    ))1(()()(

    (5)

    Aqueous phase:

    t

    SzgP

    KK wwww

    w

    rww

    )()(

    (6)

    Where 1 gw SS

    For Eq. (5) and (6), the left hand side refers to the difference between the mass flux into and out of the system. Mass

    transfer in the gaseous phase occurs either through diffusive flow or Darcy flow. However, we will neglect the diffusive term

    in the mass balance equation of the aqueous phase, because in the liquid its effect is negligible compared to that of Darcy

    flow. On the other hand, the right hand side represents the accumulation of the gas and aqueous phase. For gas we consider

    both compression (the first term) and desorption (the second term). However, we just consider the compressibility of the

    aqueous phase in the mass accumulation.

    Fig. 5Capillary curve used for the base case

    Fig. 6Relative permeability curve used for the Micro-Scale Model

    0

    1000

    2000

    3000

    4000

    5000

    0.0

    5.0

    10.0

    15.0

    20.0

    25.0

    30.0

    35.0

    40.0

    0.0 0.2 0.4 0.6 0.8 1.0

    Pc,psi

    Pc,MPa

    Water Saturation

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1

    0 0.2 0.4 0.6 0.8 1

    Kr

    Kr

    Water Saturation

    Krg_drainage

    Krg_imbibition (used in this paper)

    Krw_imbibition (used in this paper)

  • 7/25/2019 Papers Spe 166173 Ms

    6/17

    6 SPE 166173

    Table 1Hydraulic properties of the base model

    Parameter Value

    1.5

    eP , MPa 4.0

    wrS 0.2

    wcnwS 0.25

    nwcS 0.05

    wiS 0.25

    nw tS 0.25

    wcwS 0.2

    nwdn 3

    nw in 3.5

    wn 4

    In this work, a 50508 model (20,000 total grid blocks) is used to represent a single matrix bulk surrounded by a layerof fracture grids outside. The dimensions are 482 m482 m 62 m with matrix cubes of 10 m length and fracture

    aperture of 1 m. The number of the kerogen and the inorganic blocks are specified based on the TOC value. For the base

    case simulation, TOC is 9.0 wt%, which means 2625 randomly distributed kerogen grid blocks (including kerogen with

    micro- and nano-pores), 11199 inorganic blocks, and 6176 fracture blocks. Table 2provides more details about the properties

    of each pore system for the base case.

    In the micro model, the fracture pressure is constant at 8.6 MPa; the initial matrix pressure is set to be 17.2 MPa. The

    average matrix pressure changes with time as hydrocarbon flows from the matrix to fracture and water is imbibed into the

    matrix bulk. Initial saturation in different media varies based on their characteristics. Kerogen with micropores and nanopores

    (gas-wet pore system) are considered to have very high initial gas saturation (S g= 0.99). On the other hand, we consider the

    fracture media to be mostly saturated with water (Sw> 0.99). During the simulation, water imbibition occurs from the fracture

    into the matrix so that hydrocarbon gas flows out of the matrix. Initial water saturation in the inorganic matter can vary

    depending on the situation from 0.25 to 0.65 (Hill and Nelson 2000). For the base case, the initial water saturation in the

    inorganic matter is considered to be 0.30. Note that the provided mesh data, TOC and media properties (Table 2),wettabilities, capillary pressure and relative permeability correlations (Table 1), initial pressure, and initial water saturation

    values in different pore systems are used as the default in this paper, if not specified.

    Table 2Parameters for the base model

    Porosity system Fracture InorganicKerogen

    micropores nanoporesDensity, Kg/m

    3 -- 2.610

    31.3510

    3 1.410

    3

    Porosity 1.0 0.02 0.2 0.25

    Permeability, nD 84 50 50 --

    Diffusivity, m2/s -- -- 8.2110

    -5 2.0110

    -6

    DesorptionLangmuir pressure, MPa -- -- 10.342 10.342

    Langmuir volume, m3/Kg -- -- 1.638710

    -2 1.794810

    -2

    Results Analysis

    Before any simulation results are shown, we should note that the model is quite small in a magnitude of 1.010 -10m3, and

    through observation the dynamics in such a micro model start from1.010 -7second and deplete to final steady state after 0.1

    second. Therefore, to make the simulation results readable in a time-scale, a dimensionless time is defined as a ratio of the

    simulation time to the minimum time step size (1.010-7). Additionally, grid pressure (pressure of the non-wetting phase) is

    always plotted for different media.

    For the base case with the default parameters in the model, the results are shown in Fig. 7, Fig. 8, and Fig. 9. The

    dynamics in the inorganic matter are shown inFig. 7.In the inorganic matter, water saturation increases from the initial level

    0.30 to the final level 0.74. In the inorganic matter, gas is the non-wetting phase. Due to the initial high capillary pressure in

    the inorganic matter (Fig. 5), water is imbibed but gas is drained out. In consequence, the inorganic matter is downstream for

    water flow but upstream for gas flow a counter-current imbibition process. Reviewing the relative permeability curve in

    Fig. 6,in the Swrange of 0.3 to 0.42, the relative permeability of gas (imbibition) is very high but that of water is close to

  • 7/25/2019 Papers Spe 166173 Ms

    7/17

    SPE 166173 7

    zero; therefore, during this period, the water flowing in to the inorganic matter is limited to fracture (upstream)-inorganic

    connections at the outer part of shale matrix (Fig. 10(b)), however, the gas flowing out from the inorganic matter is supported

    by fracture-inorganic, inorganic-inorganic, and inorganic-organic (single phase) connections. The combined effect of gas

    flow-out and water flow-in is that pressure in the inorganic matter decreases. After Swbecomes greater than 0.42 (Fig. 6), the

    inorganic-inorganic connection also joins in the water flow process so that water can easily get into the inner shale matrix.

    When Swin the inorganic matter reaches to 0.64, gas relative permeability in the inorganic matter is close to zero but that of

    water keeps increasing. In consequence, in the inorganic matter, water flow-in is enhanced but gas is trapped there.

    Therefore, when Swis greater than about 0.64, pressure in the inorganic matter increases and finally reaches to a steady level(Fig. 7). On the other hand, through the Swchange in the inorganic matter, the water imbibition into the inorganic matrix is

    extremely active with an increase of about 0.44 pore volume (that of inorganic pore system) of water imbibed.

    Fig. 7Dynamics in the inorganic matter including average Swand average pressure

    Fig. 8Dynamics in the kerogen including average Swand average pressure

    0.0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1.0

    0.0

    2.0

    4.0

    6.0

    8.0

    10.0

    12.0

    14.0

    16.0

    18.0

    20.0

    1.0E+00 1.0E+01 1.0E+02 1.0E+03 1.0E+04 1.0E+05 1.0E+06

    WaterSaturation

    Pressure,MPa

    Dimensionless Time

    Gas Pressure in the Inorganic Matter

    Sw in the Inorganic Matter

    0.0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1.0

    0.0

    2.0

    4.0

    6.0

    8.0

    10.0

    12.0

    14.0

    16.0

    18.0

    20.0

    1.0E+00 1.0E+01 1.0E+02 1.0E+03 1.0E+04 1.0E+05 1.0E+06

    WaterSaturation

    Pressure,MPa

    Dimensionless Time

    Pressure in the Kerogen

    Sw in the Kerogen

  • 7/25/2019 Papers Spe 166173 Ms

    8/17

    8 SPE 166173

    Fig. 9Gas recovery from shale matrix to surrounding fractures

    The dynamics in the kerogen are shown in Fig. 8. Kerogen is gas-wetting and upstream for water flow. With Sw

    approximately constant at 0.01 in the kerogen (Fig. 8), water cannot flow out from the kerogen and consequently only singlegas phase flows there. Therefore, pressure change in the kerogen is only caused by gas flow through the kerogen. At the early

    production period, gas flows from the inorganic matter into the kerogen, and simultaneously gas also flows from the kerogen

    into the fracture system. These two gas flow pathways have the opposite effect on average pressure change in the kerogen.

    Initially they compensate each other and thus an approximately constant level of pressure at the early production period is

    observed in Fig. 8. Since water continues to imbibe into the inorganic matter, in the inorganic matter water saturation

    increases and gas relative permeability decreases. Therefore, at a later production period, gas flowing from the inorganic

    matter to the kerogen gradually becomes weak but gas keeps flowing out from the kerogen to the fractures, and thus the

    pressure decreases in the kerogen. Finally, a steady state in the kerogen is reached when the micro model is in equilibrium.

    The gas recovery from the shale matrix into the surrounding fracture system is plotted inFig. 9.The ultimate gas recovery

    is about 30%. In this case, there are 6 horizontal layers of matrix in the micro model. Here the second layer of matrix is taken

    as an example, and the grid map is shown inFig. 11.Those yellow grids are the inorganic grids, and the blue crosses here are

    kerogen with micropores and kerogen with nanopores, and the brown grids surrounding the matrix square are the fracture

    grids. The kerogen grids are randomly distributed within this layer.Fig. 10 shows grid pressure and water saturation changesin this layer in a time sequence. InFig. 10(a), the pressure in the inorganic matter drops because it is predominated by the gas

    release, yet the water saturation in the inorganics exhibits little change; in the kerogen the pressure stays high and S wthere is

    at 0.01.Fig. 10(b) shows the dynamics afterFig. 10(a). The pressure in the inorganics mostly further decrease, and S win the

    inorganic matter neighboring fractures increases; in the kerogen, the pressure decreases moderately and water saturation is

    still 0.01. Fig. 10(c) is the final time step of the simulation. S w in the inorganic matter increases to a high level and the

    pressure there also builds up and is higher than that in the kerogen.

    Sensitivity Analysis

    In this part, the effects of desorption, diffusion, TOC, initial Sw and capillary pressure curve in the inorganic matter are

    analyzed. Those corresponding parameters are sensitized here and any other inputs for the simulation are the same as that in

    the base case in the last section. To better compare with the base case, cumulative gas production in all the cases below is

    normalized by the gas in place in the base case, namely normalized cumulative gas production.

    Effect of Desorption

    The effect of desorption is analyzed by changing the Langmuir desorption parameters in the kerogen (Table 3), with other

    inputs for the simulation the same as the default. Here five cases with an increase of adsorption gas in the kerogen are

    analyzed. The results for each case are plotted inFig. 12 andFig. 13.Note that in the plot the average Langmuir parameters

    are shown. InFig. 12 andFig. 13,with an increase in the Langmuir desorption in the kerogen, the gas production increases

    correspondingly. However, here there is not much significant increase. We consider that gas production is determined by the

    Langmuir curves and the pressure drop in the kerogen (Fig. 13). InFig. 13, the pressure drops in the kerogen for the five

    cases are almost the same. The average pressure in the kerogen is always greater than the Langmuir pressure in the kerogen

    (10.342 MPa). Because the Langmuir pressure is the pressure corresponding to half of the Langmuir volume, in all of those

    cases the desorbed gas in the kerogen is less than half of the Langmuir volume. Therefore, only small increases in gas

    production can be expected after desorption is introduced.

    0.0

    5.0

    10.0

    15.0

    20.0

    25.0

    30.0

    35.0

    0.0

    5.0

    10.0

    15.0

    20.0

    25.0

    30.0

    35.0

    0.0E+00 5.0E+04 1.0E+05 1.5E+05 2.0E+05 2.5E+05 3.0E+05 3.5E+05 4.0E+05

    GasRecovery,%

    GasRecovery,%

    Dimensionless Time

  • 7/25/2019 Papers Spe 166173 Ms

    9/17

    SPE 166173 9

    (a) Early time

    (b) Later time

    (c) Final time

    Fig. 10Pressure and water saturation in different time steps

    Fig. 11Grid map for the matrix layer, blue crosses are the kerogen grids; yellow grids are the inorganic grids; brown grids are thefracture grids.

  • 7/25/2019 Papers Spe 166173 Ms

    10/17

    10 SPE 166173

    Table 3Parameters for the desorption analysis (Case 4-Base Case)

    Case NO.

    Kerogen with micropores Kerogen with nanopores

    Langmuirpressure, MPa

    Langmuirvolume, m

    3/Kg

    Langmuirpressure, MPa

    Langmuirvolume, m

    3/Kg

    Case 1 NA NA NA NA

    Case 2 10.342 1.041510-2 10.342 1.170510

    -2

    Case 3 10.342 1.326610-2 10.342 1.482710

    -2

    Case 4 10.342 1.638710-2 10.342 1.794810

    -2

    Case 5 10.342 1.950910-2 10.342 2.106910

    -2

    Fig. 12Gas production for cases with different Langmuir parameters in the kerogen

    Fig. 13

    Average pressure in the kerogen for cases with different Langmuir parameters

    Effect of Diffusion

    Diffusion occurs in the kerogen in the micro model. Through changing the diffusion coefficient shown in Table 4, the

    effect of diffusion is also analyzed with the other inputs for the simulation the same as the default. Here five cases with an

    increase of the diffusion coefficient in the kerogen are analyzed. The results are plotted inFig. 14 andFig. 15.InFig. 14,the

    normalized cumulative gas production in Case 1 is far lower than other cases (2 to 5), and the normalized cumulative gas

    productions in Case 2 to 5 are all the same (30.0%). Since diffusion is deactivated in Case 1, gas production in Case 1 is only

    through Darcy flow. As diffusion occurs only in the kerogen, the gap of cumulative production between Case 1 and other

    cases implies the importance of kerogen to the gas production process. In addition to that, lower gas production in Case 1

    shows the indispensable role of diffusion, besides Darcy flow, as a main production mechanism in the kerogen network. The

    separation of Case 1 and other cases after the early production period in Fig. 14 also indicates that diffusion occurs following

    0.0

    5.0

    10.0

    15.0

    20.0

    25.0

    30.0

    35.0

    40.0

    0.0

    5.0

    10.0

    15.0

    20.0

    25.0

    30.0

    35.0

    40.0

    1.0E+00 1.0E+01 1.0E+02 1.0E+03 1.0E+04 1.0E+05 1.0E+06

    NormalizedCumulativeGasProduction,%

    NormalizedCumulativeGasProduction,%

    Dimensionless Time

    Case 1-No Desorption

    Case 2-VL = 1.09E-2 m/Kg, PL = 10.34 MPa

    Case 3-VL = 1.40E-2 m/Kg, PL = 10.34 MPa

    Case 4-VL = 1.72E-2 m/Kg, PL = 10.34 MPa

    Case 5-VL = 2.03E-2 m/Kg, PL = 10.34 MPa

    0.0

    2.0

    4.0

    6.0

    8.0

    10.0

    12.0

    14.0

    16.0

    18.0

    20.0

    0.0

    2.0

    4.0

    6.0

    8.0

    10.0

    12.0

    14.0

    16.0

    18.0

    20.0

    1.0E+00 1.0E+01 1.0E+02 1.0E+03 1.0E+04 1.0E+05 1.0E+06

    Pressure,MPa

    Pressure,MPa

    Dimensionless Time

    Case 1-No Desorption

    Case 2-VL = 1.09E-2 m/Kg, PL = 10.34 MPa

    Case 3-VL = 1.40E-2 m/Kg, PL = 10.34 MPa

    Case 4-VL = 1.72E-2 m/Kg, PL = 10.34 MPa

    Case 5-VL = 2.03E-2 m/Kg, PL = 10.34 MPa

  • 7/25/2019 Papers Spe 166173 Ms

    11/17

    SPE 166173 11

    Darcy flow. Further, with diffusion considered in Case 2 to 5, the ultimate cumulative gas production is almost the same.

    This implies that different diffusion coefficients can actually accelerate the gas production process, but do not change the

    ultimate production. Fig. 15 shows the average pressure in the kerogen. In Case 1, pressure changes little in the kerogen

    because diffusion is not considered. In Cases 2 to 5, diffusion is considered and it is clearly helpful in accelerating the

    pressure depletion in the kerogen.

    Table 4Parameters for the diffusion analysis (Case 4-Base Case)

    Case NO. Diffusion Coefficientin Kerogen with micropores, m2/s Diffusion Coefficientin Kerogen with nanopores, m

    2/s

    Case 1 NA NA

    Case 2 8.2110-7 2.0110

    -8

    Case 3 8.2110-6 2.0110

    -7

    Case 4 8.2110-5 2.0110

    -6

    Case 5 8.2110-4 2.0110

    -5

    Fig. 14Gas production in the kerogen for cases with different diffusion coefficient

    Fig. 15Average pressure in the kerogen for cases with different diffusion coefficient

    Effect of TOC

    TOC (wt %) is a necessary parameter to the discretized the micro model because it controls the number of kerogen grids

    (Yan et al. 2013b; Yan et al. 2013a). Here four TOC values are compared, 3.0 wt%, 6.0 wt%, 9.0 wt% (base case), 12.0 wt%.

    Therefore, four mesh maps are generated with different numbers of kerogen grids and input into the simulator. The results are

    0.0

    5.0

    10.0

    15.0

    20.0

    25.0

    30.0

    35.0

    40.0

    0.0

    5.0

    10.0

    15.0

    20.0

    25.0

    30.0

    35.0

    40.0

    1.0E+00 1.0E+01 1.0E+02 1.0E+03 1.0E+04 1.0E+05 1.0E+06

    NormalizedCumulativeGas

    Production,%

    NormalizedCumulativeGas

    Production,%

    Dimensionless Time

    Case 1-No DiffusionCase 2-D_micro = 8.21E-7 m/s, D_nano = 2.01E-8 m/s

    Case 3-D_micro = 8.21E-6 m/s, D_nano = 2.01E-7 m/sCase 4-D_micro = 8.21E-5 m/s, D_nano = 2.01E-6 m/s

    Case 5-D_micro = 8.21E-4 m/s, D_nano = 2.01E-5 m/s

    0.0

    2.0

    4.0

    6.0

    8.0

    10.0

    12.0

    14.0

    16.0

    18.0

    20.0

    0.0

    2.0

    4.0

    6.0

    8.0

    10.0

    12.0

    14.0

    16.0

    18.0

    20.0

    1.0E+00 1.0E+01 1.0E+02 1.0E+03 1.0E+04 1.0E+05 1.0E+06

    Pressure,MPa

    Pressure,MPa

    Dimensionless Time

    Case 1- No Diffusion

    Case 2-D_micro = 8.21E-7 m/s, D_nano = 2.01E-8 m/s

    Case 3-D_micro = 8.21E-6 m/s, D_nano = 2.01E-7 m/s

    Case 4-D_micro = 8.21E-5 m/s, D_nano = 2.01E-6 m/s

    Case 5-D_micro = 8.21E-4 m/s, D_nano = 2.01E-5 m/s

  • 7/25/2019 Papers Spe 166173 Ms

    12/17

    12 SPE 166173

    plotted inFig. 16.The figure shows that the higher the TOC, the greater the ultimate gas cumulative production. TOC is the

    weight percentage of organic carbon (kerogen) in shale matrix. In the kerogen in the micro model, desorption occurs and

    there is also considerable free gas. Therefore, higher TOC content means more kerogen in the shale matrix and thus more gas

    in place provided for gas production (Yan et al. 2013a).Fig. 16 clearly shows that the higher the TOC, the higher the ultimate

    gas cumulative production. In addition, the slopes of gas cumulative production curves are actually determined by the

    production rate. Fig. 16 shows that the higher the TOC value, the larger the slope of the curve of the gas cumulative

    production during the main production period because of an increasing effect of diffusion.

    Fig. 16Gas production for cases with TOC value in the shale matrix

    Fig. 17Gas production for cases with different initial Swin the inorganic matter

    Effect of Initial Water Saturation in the Inorganic Matter

    In the micro model, the inorganic matter is water wet, and gas in this pore system is easily blocked by water, shown as Krg

    imbibition curve inFig. 6.Therefore, initial water saturation (Sw) may influence the gas production and water imbibition in

    the inorganic matter. Here five initial Sw values in the inorganic matter are sensitized, 0.21 (slightly higher than the

    irreducible water saturation 0.20), 0.3 (base case), 0.4, 0.5, and 0.6. The results are plotted in Fig. 17 andFig. 18.Higher

    initial Swhere means lower gas saturation in the inorganic matter. Therefore, with the same initial pressure and temperature,

    cases with higher initial water saturation should have moderately lower gas in place even though porosity in the inorganic

    matter is relatively low. Therefore, we observe that higher initial Swin the inorganic matter brings lower gas production in the

    shale matrix (Fig. 17). InFig. 18,the average Swin the inorganic matter is plotted. Those cases with higher initial S whave

    less water imbibition capacity and all of them finally reach a final saturation of about 0.73.

    0.0

    5.0

    10.0

    15.0

    20.0

    25.0

    30.0

    35.0

    40.0

    0.0

    5.0

    10.0

    15.0

    20.0

    25.0

    30.0

    35.0

    40.0

    1.0E+00 1.0E+01 1.0E+02 1.0E+03 1.0E+04 1.0E+05 1.0E+06

    NormalizedCumulativeGasProduction,%

    Norm

    alizedCumulativeGasProduction,%

    Dimensionless Time

    Case 1- TOC = 3.0 wt%

    Case 2- TOC = 6.0 wt%

    Case 3- TOC = 9.0 wt%

    Case 4- TOC = 12.0 wt%

    0.0

    5.0

    10.0

    15.0

    20.0

    25.0

    30.0

    35.0

    40.0

    0.0

    5.0

    10.0

    15.0

    20.0

    25.0

    30.0

    35.0

    40.0

    1.0E+00 1.0E+01 1.0E+02 1.0E+03 1.0E+04 1.0E+05 1.0E+06

    NormalizedCumulativeGasProduction,%

    NormalizedCumulativeGasProduction,%

    Dimensionless Time

    Case 1- Sw = 0.21Case 2- Sw = 0.30Case 3- Sw = 0.40Case 4- Sw = 0.50Case 5- Sw =0.60

  • 7/25/2019 Papers Spe 166173 Ms

    13/17

    SPE 166173 13

    Fig. 18Average Swin the inorganic matter for cases with different initial Swin the inorganic matter

    Effect of Capillary Pressure Curve in the Inorganic MatterTo better investigate the effect of capillary forces, as an important factor during spontaneous imbibition, we have

    analyzed three different capillary pressure scenarios (Fig. 19). Case 1 represents a system with fairly even distribution of the

    pore sizes and relatively small entry pressure, which serves as the low capillary pressure case here. Case 2 (the base case),

    has higher capillary entry pressure but same pore size distribution. Case 3, on the other hand, has a slightly lower pore size

    distribution factor (pore sizes are distributed more widely) with a higher entry pressure that results in higher capillary

    pressure values (capillary pressures in the inorganic blocks for cases 1 to 3 are 6, 16 and 37 MPa respectively, at the initial Sw

    of 0.3). To better analyze the imbibition process and see the effect of inorganic network on gas production, capillary pressure

    parameters in the organic grids will be considered constant and the changes will be imposed solely upon the inorganic grids.

    Looking at the average water saturation in the inorganic grid blocks (Fig. 20), it is apparent that higher capillary pressures

    accelerate the imbibition process. In addition to that, average water saturation at the late times is slightly higher in high

    capillary cases due to an improved water influx. On the other hand, when we look at the gas production graph (Fig. 21), we

    see capillary pressure (in the inorganic network) does not have considerable effect on ultimate gas production. This can be

    interpreted as the important effect of kerogen grid blocks and different production mechanisms there (i.e. diffusion) on gasproduction as it was also observed in the diffusion sensitivity analysis.

    Fig. 19Three different capillary pressure curves for the inorganic matter

    0.0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1.0

    0.0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1.0

    1.0E+00 1.0E+01 1.0E+02 1.0E+03 1.0E+04 1.0E+05 1.0E+06

    WaterSatu

    ration

    WaterSatu

    ration

    Dimensionless Time

    Case 1- Sw = 0.21 Case 2- Sw = 0.30

    Case 3- Sw = 0.40 Case 4- Sw = 0.50

    Case 5- Sw = 0.60

    0

    1000

    2000

    3000

    4000

    5000

    0.0

    5.0

    10.0

    15.0

    20.0

    25.0

    30.0

    35.0

    40.0

    0.0 0.2 0.4 0.6 0.8 1.0

    Pc,psi

    Pc,MPa

    Water Saturation

    Case 1 - = 1.5; Entry Pressure: 1.5 MPa

    Case 2 - = 1.5; Entry Pressure: 4.0 MPa

    Case 3 - = 1.15; Entry Pressure: 6.0 MPa

  • 7/25/2019 Papers Spe 166173 Ms

    14/17

    14 SPE 166173

    Fig. 20Average Swin the inorganic matter for cases with different capillary pressure in the inorganic matter

    Fig. 21Gas production for cases with different capillary pressure in the inorganic matter

    Conclusions

    A Micro-Scale Model considering multiple pore systems, different mechanisms and mixed wettability has beenestablished. This model allows the interpretation of different complex flow interactions in shale.

    The dynamics in the Micro-Scale Model are complex, and might be controlled by multiple factors at the same time.

    The capacity of water imbibition into the inorganic matter is considerable. Under high water saturation in theinorganic matter, gas is trapped there, and further water imbibition makes pressure in the inorganic matter increase.

    Desorption can increase the gas in place in shale but its effect on gas production rates is not significant because ofthe limited pressure drop in the kerogen.

    Diffusion, as a gas-phase transport mechanism in the kerogen, plays an important role in total gas production byfacilitating the gas flow from the hydrocarbon-rich organic matter to the fracture network.

    TOC content in the kerogen can increase the gas production with a combined effect of diffusion, desorption and freegas storage.

    Higher capillary pressure in the inorganic matter accelerates the water imbibition process but has little influence ongas production.

    0.0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1.0

    0.0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1.0

    1.0E+00 1.0E+01 1.0E+02 1.0E+03 1.0E+04 1.0E+05 1.0E+06

    WaterSatu

    ration

    WaterSatu

    ration

    Dimensionless Time

    Case 1 - = 1.5; Entry Pressure: 1.5 MPa

    Case 2 - = 1.5; Entry Pressure: 4.0 MPa

    Case 3 - = 1.15; Entry Pressure: 6.0 MPa

    0.0

    5.0

    10.0

    15.0

    20.0

    25.0

    30.0

    35.0

    40.0

    0.0

    5.0

    10.0

    15.0

    20.0

    25.0

    30.0

    35.0

    40.0

    1.0E+00 1.0E+01 1.0E+02 1.0E+03 1.0E+04 1.0E+05 1.0E+06

    NormalizedCumulativeGasProduction,%

    NormalizedCumulativeGasProduction,%

    Dimensionless Time

    Case 1 - = 1.5; Entry Pressure: 1.5 MPaCase 2 - = 1.5; Entry Pressure: 4.0 Mpa

    Case 3 - = 1.15; Entry Pressure: 6.0 Mpa

  • 7/25/2019 Papers Spe 166173 Ms

    15/17

    SPE 166173 15

    Nomenclature

    gC = gas compressibility, 1/Pa;

    D = diffusion coefficient, m2/second;g = the gravitational acceleration vector, m2/s;

    K = media permeability, m2;

    rgK = gas relative permeability, dimensionless;

    rnwdK = non-wetting phase drainage relative permeability, dimensionless;

    rnwiK = non-wetting phase imbibition relative permeability, dimensionless;

    rwK = water relative permeability, or wetting phase drainage/imbibition relative permeability, dimensionless;

    nwdn = non-wetting phase drainage exponent, dimensionless;

    nw in = non-wetting phase imbibition exponent, dimensionless;

    wn = wetting phase drainage/imbibition exponent, dimensionless;

    P = grid pressure, Pa;

    cP = capillary pressure, Pa;

    eP = capillary entry pressure, Pa;

    gP = gas phase pressure, Pa;

    wP = water phase pressure, Pa;

    aq = the mass of gas adsorbed on unit volume of media, kg/m3;

    eS = effective wetting phase saturation, fraction;

    gS = gas saturation, fraction;

    nwcS = non-wetting phase critical saturation, fraction;

    nwtS = non-wetting phase trapped saturation, fraction;

    wS = water saturation, or wetting phase saturation, fraction;

    wiS = initial wetting phase saturation, fraction;

    wcnwS = critical wetting phase saturation with respect to the non-wetting phase, fraction;

    wcwS = critical wetting phase saturation with respect to the wetting phase, fraction;

    wrS = wetting phase residual saturation, fraction;

    t = time, seconds;

    TOC = weight percentage of Total Organic Carbon, wt%z = distance in the gravitational direction, m;

    g = gas density at reservoir conditions, kg/m3;

    w = water density at reservoir conditions, kg/m3;

    g = gas viscosity, Pas;

    w = water viscosity, Pas;

    = the porosity of the porous media, fraction;

    = pore size distribution index, fraction;

    Acknowledgements

    The authors wish to thank The Crisman Institute at Texas A&M University for funding this project.

  • 7/25/2019 Papers Spe 166173 Ms

    16/17

    16 SPE 166173

    References

    Brooks, R.H. and Corey, A.T. 1964. Hydraulic Properties of Porous Media. Colorado State University, Hydrology Papers

    (5).

    Civan, F., Rai, C., and Sondergeld, C. 2011. Shale-Gas Permeability and Diffusivity Inferred by Improved Formulation of

    Relevant Retention and Transport Mechanisms. Transport in Porous Media86(3): 925-944. DOI: 10.1007/s11242-

    010-9665-x

    Cui, X., Bustin, A.M.M., and Bustin, R.M. 2009. Measurements of Gas Permeability and Diffusivity of Tight Reservoir

    Rocks: Different Approaches and Their Applications. Geofluids 9 (3): 208-223. DOI: 10.1111/j.1468-8123.2009.00244.x

    Curtis, M.E., Ambrose, R.J., and Sondergeld, C.H. 2010. Structural Characterization of Gas Shales on the Micro- and Nano-

    Scales. Paper presented at the Canadian Unconventional Resources and International Petroleum Conference,

    Calgary, Alberta, Canada. Society of Petroleum Engineers SPE-137693-MS. DOI: 10.2118/137693-ms.

    Dacy, J.M. 2010. Core Tests for Relative Permeability of Unconventional Gas Reservoirs. Paper presented at the SPE Annual

    Technical Conference and Exhibition, Florence, Italy. Society of Petroleum Engineers SPE-135427-MS. DOI:

    10.2118/135427-ms.

    Darishchev, A., de Nancy, E.N.S.G., Lemouzy, P. et al. 2013. On Simulation of Flow in Tight and Shale Gas Reservoirs.

    Paper presented at the 2013 SPE Middle East Unconventional Gas Conference & Exhibition, Muscat, Sultanate of

    Oman. Society of Petroleum Engineers SPE-163990-MS. DOI: 10.2118/163990-ms.

    Ertekin, T., King, G.A., and Schwerer, F.C. 1986. Dynamic Gas Slippage: A Unique Dual-Mechanism Approach to the Flow

    of Gas in Tight Formations. SPE Formation Evaluation1(1): 43-52. DOI: 10.2118/12045-pa

    Freeman, C., Moridis, G.J., Michael, G.E. et al. 2012. Measurement, Modeling, and Diagnostics of Flowing Gas CompositionChanges in Shale Gas Wells. Paper presented at the SPE Latin America and Caribbean Petroleum Engineering

    Conference, Mexico City, Mexico. Society of Petroleum Engineers SPE-153391-MS. DOI: 10.2118/153391-ms.

    Hill, D.G. and Nelson, C.R. 2000. Gas Productive Fractured Shales an Overview and Update. Gas TIPS, 4-13.

    Hu, Y., Devegowda, D., Striolo, A. et al. 2013. A Pore Scale Study Describing the Dynamics of Slickwater Distribution in

    Shale Gas Formations Following Hydraulic Fracturing. Paper presented at the SPE Unconventional Resources

    Conference - USA, The Woodlands, TX, USA. 2013, Society of Petroleum Engineers SPE-164552-MS. DOI:

    10.2118/164552-ms.

    InGrain. 2010. Nano 3 Image Slice Distinguishes Kerogen from the Pore Space, http://www.ingrainrocks.com/

    Javadpour, F. 2009. Nanopores and Apparent Permeability of Gas Flow in Mudrocks (Shales and Siltstone). Journal of

    Canadian Petroleum Technology48(8): 16-21. DOI: 10.2118/09-08-16-da

    Javadpour, F., Fisher, D., and Unsworth, M. 2007. Nanoscale Gas Flow in Shale Gas Sediments. Journal of Canadian

    Petroleum Technology46(10). DOI: 10.2118/07-10-06

    Ji, L., Zhang, T., Milliken, K.L. et al. 2012. Experimental Investigation of Main Controls to Methane Adsorption in Clay-Rich Rocks.Applied Geochemistry27(12): 2533-2545. DOI:http://dx.doi.org/10.1016/j.apgeochem.2012.08.027

    Kalakkadu, S., Alba, A.K., Callard, J. et al. 2013. Production Analysis in Multifrac Horizontal Wells Including the Effects of

    Adsorption, Mixed Wettability and Dual Porosity. Paper presented at the 2013 SPE Production and Operations

    Symposium, Oklahoma City, Oklahoma, USA. Society of Petroleum Engineers SPE-164479-MS. DOI:

    10.2118/164479-ms.

    King, G.E. 2012. Hydraulic Fracturing 101: What Every Representative, Environmentalist, Regulator, Reporter, Investor,

    University Researcher, Neighbor and Engineer Should Know About Estimating Frac Risk and Improving Frac

    Performance in Unconventional Gas and Oil Wells. Paper presented at the SPE Hydraulic Fracturing Technology

    Conference, The Woodlands, Texas, USA. Society of Petroleum Engineers SPE-152596-MS. DOI:

    10.2118/152596-ms.

    Molina, N.N. 1980. A Systematic Approach to the Relative Permeability Problem in Reservoir Simulation. Paper presented at

    the SPE Annual Technical Conference and Exhibition, Dallas, Texas. 1980 Copyright 1980, American Institute of

    Mining, Metallurgical, and Petroleum Engineers, Inc. 00009234. DOI: 10.2118/9234-ms.

    Passey, Q.R., Bohacs, K., Esch, W.L. et al. 2010. From Oil-Prone Source Rock to Gas-Producing Shale Reservoir Geologicand Petrophysical Characterization of Unconventional Shale-Gas Reservoirs. Paper presented at the International Oil

    and Gas Conference and Exhibition in China, Beijing, China. Society of Petroleum Engineers SPE-131350-MS.

    DOI: 10.2118/131350-ms.

    Shabro, V., Torres-Verdin, C., and Javadpour, F. 2011. Numerical Simulation of Shale-Gas Production: From Pore-Scale

    Modeling of Slip-Flow, Knudsen Diffusion, and Langmuir Desorption to Reservoir Modeling of Compressible

    Fluid. Paper presented at the North American Unconventional Gas Conference and Exhibition, The Woodlands,

    Texas, USA. Society of Petroleum Engineers SPE-144355-MS. DOI: 10.2118/144355-ms.

    Shabro, V., Torres-Verdin, C., and Sepehrnoori, K. 2012. Forecasting Gas Production in Organic Shale with the Combined

    Numerical Simulation of Gas Diffusion in Kerogen, Langmuir Desorption from Kerogen Surfaces, and Advection in

    Nanopores. Paper presented at the SPE Annual Technical Conference and Exhibition, San Antonio, Texas, USA.

    Society of Petroleum Engineers SPE-159250-MS. DOI: 10.2118/159250-ms.

    http://dx.doi.org/10.1016/j.apgeochem.2012.08.027http://dx.doi.org/10.1016/j.apgeochem.2012.08.027
  • 7/25/2019 Papers Spe 166173 Ms

    17/17

    SPE 166173 17

    Swami, V. and Settari, A. 2012. A Pore Scale Gas Flow Model for Shale Gas Reservoir. Paper presented at the SPE

    Americas Unconventional Resources Conference, Pittsburgh, Pennsylvania USA. Society of Petroleum Engineers

    SPE-155756-MS. DOI: 10.2118/155756-ms.

    Wang, F.P. and Reed, R.M. 2009. Pore Networks and Fluid Flow in Gas Shales. Paper presented at the SPE Annual

    Technical Conference and Exhibition, New Orleans, Louisiana. Society of Petroleum Engineers SPE-124253-MS.

    DOI: 10.2118/124253-ms.

    Wang, Y., Yan, B., and Killough, J.E. 2013. Compositional Modeling of Tight Oil Using Dynamic Nanopore Properties.

    Paper presented at the 2013 SPE Annual Technical Conference and Exhibition, New Orleans, Louisiana. Society ofPetroleum Engineers SPE 166267.

    Wei, C., Qin, G., Guo, W. et al. 2013. Characterization and Analysis on Petrophysical Parameters of a Marine Shale Gas

    Reservoir. Paper presented at the SPE Western Regional & AAPG Pacific Section Meeting, 2013 Joint Technical

    Conference, Monterey, CA, USA. Society of Petroleum Engineers SPE-165380-MS. DOI: 10.2118/165380-ms.

    Yan, B. 2013. A Novel Approach for the Simulation of Multiple Flow Mechanisms and Porosities in Shale Gas Reservoirs.

    M.S. Thesis, Texas A&M University, College Station.

    Yan, B., Killough, J., Wang, Y. et al. 2013b. Novel Approaches for the Simulation of Unconventional Reservoirs. Paper

    presented at the Unconventional Resources Technology Conference, Denver, Colorado, USA. SPE, AAPG, SEG

    URTeC 1581172.

    Yan, B., Wang, Y., and Killough, J. 2013a. Beyond Dual-Porosity Modeling for the Simulation of Complex Flow

    Mechanisms in Shale Reservoirs. Paper presented at the 2013 SPE Reservoir Simulation Symposium, The

    Woodlands, TX, USA. Society of Petroleum Engineers SPE-163651-MS. DOI: 10.2118/163651-ms.

    Zhang, T., Ellis, G.S., Ruppel, S.C. et al. 2012. Effect of Organic-Matter Type and Thermal Maturity on Methane Adsorption

    in Shale-Gas Systems. Organic Geochemistry 47 (0): 120-131. DOI:

    http://dx.doi.org/10.1016/j.orggeochem.2012.03.012

    http://dx.doi.org/10.1016/j.orggeochem.2012.03.012http://dx.doi.org/10.1016/j.orggeochem.2012.03.012