1-s2.0-s187551001400314x-main

Upload: kilaparthi-satyavamma

Post on 01-Jun-2018

215 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/9/2019 1-s2.0-S187551001400314X-main

    1/12

    Conceptual study of thermal stimulation in shale gas formations

    HanYi Wang  a ,  *, Omobola Ajao  b , Michael J. Economides  a , 1

    a University of Houston, United Statesb EPT International, United States

    a r t i c l e i n f o

     Article history:

    Received 8 September 2014

    Received in revised form

    11 October 2014

    Accepted 13 October 2014

    Available online 30 October 2014

    Keywords:

    Shale gas

    Adsorption

    Desorption

    Thermal recovery

    Hydraulic fracture

    Stimulation

    a b s t r a c t

    Shale gas formations have become a major source of energy in recent years. Developments in hydraulic

    fracturing technology have made these reservoirs more accessible and productive. Apart from otherdissimilarities from conventional gas reservoirs, one major difference is that a considerable amount of 

    gas produced from these shale gas formations comes from desorption. Up to 85% of the total gas within

    shale can be found as an adsorbed phase on clay and kerogen, so how much adsorbed gas can be pro-

    duced will have signicant impact on ultimate gas recovery. The Langmuir isotherm has been widely

    used in industry to describe the pressure dependence of adsorbed gas. However, temperature dependent

    adsorption behavior and its major implications for evaluating thermal stimulation as a recovery method

    for shale reservoirs have not been thoroughly explored. Therefore, in order to design and analyze the

    thermal treatment of shale gas formations successfully, it is crucial to understand the effects of fracture

    heating on the shale gas adsorption and desorption phenomenon, and how can we exploit such effects to

    enhance shale gas recovery from hydraulically fractured reservoirs. Even though numerous research

    efforts have been focused on thermal recovery of shale oil, its possible application to shale gas has not

    been investigated.

    In this research, we propose a method to evaluate desorbed gas as a function of pressure and tem-

    perature in shale formations, by regression of a Bi-Langmuir model on Langmuir isotherm data. We have

    developed a fully coupled unconventional reservoir simulator, which is capable of capturing real gas  ow

    in the shale matrix and in the hydraulic fracture by accounting for the effects of gas desorption anddiffusion, as well as the temperature diffusion process within the matrix. This simulator enables us to

    investigate the effects of fracture heating on the shale gas desorption phenomenon on the global well

    performance and recovery. The results of this study show, for the  rst time in a rigorous way, that by

    increasing the temperature within the fracture, shale gas recovery can be improved.

    We have rationalized and quantied relations between the adsorption/desorption fundamental phe-

    nomena and stimulation temperature, fracture spacing, reservoir permeability and bottom hole pressure.

    The thermal properties of shale formations only have limited impacts on long term production. The

    results of this study can provide a guidance to develop a strategy to design thermal treatment in hy-

    draulically fractured shale formations and propose the degree of thermal stimulation temperature

    required in a fracture to promote an economically viable return on production.

    ©  2014 Elsevier B.V. All rights reserved.

    1. Introduction

    Unconventional gas reservoirs are expectedto play a vital role in

    satisfying the global demand for gas in the future. The major

    component of unconventional gas reservoirs comprises of shale

    gas. Shales and silts are the most abundant sedimentary rocks in

    the earth's crust and it is evident from the recent year's activities in

    shale gas plays that in the future shale gas will constitute the largest

    component in gas production globally. Unlike conventional gas

    reservoirs, shale gas reservoirs have very low permeability, and are

    economical only when hydraulically fractured. The key techniques

    that allow extracting shale gas commercially such as horizontal

    drilling and hydraulic fracturing, are expected to improve with

    time; however as better stimulation techniques are becoming

    attainable, it is important to have better understanding of shale gas

    reservoir behavior in order to apply these techniques in an ef cient

    fashion. One important aspect of shale gas reservoirs which needs

    special consideration is the adsorption/desorption phenomenon.*   Corresponding author.

    E-mail address: [email protected] (H. Wang).1 Deceased.

    Contents lists available at  ScienceDirect

     Journal of Natural Gas Science and Engineering

    j o u r n a l h o m e p a g e :   w w w . e l s e v i e r . c om / l o c a t e / j n g s e

    http://dx.doi.org/10.1016/j.jngse.2014.10.015

    1875-5100/©

     2014 Elsevier B.V. All rights reserved.

     Journal of Natural Gas Science and Engineering 21 (2014) 874e885

    mailto:[email protected]://www.sciencedirect.com/science/journal/18755100http://www.elsevier.com/locate/jngsehttp://dx.doi.org/10.1016/j.jngse.2014.10.015http://dx.doi.org/10.1016/j.jngse.2014.10.015http://dx.doi.org/10.1016/j.jngse.2014.10.015http://dx.doi.org/10.1016/j.jngse.2014.10.015http://dx.doi.org/10.1016/j.jngse.2014.10.015http://dx.doi.org/10.1016/j.jngse.2014.10.015http://www.elsevier.com/locate/jngsehttp://www.sciencedirect.com/science/journal/18755100http://crossmark.crossref.org/dialog/?doi=10.1016/j.jngse.2014.10.015&domain=pdfmailto:[email protected]

  • 8/9/2019 1-s2.0-S187551001400314X-main

    2/12

    In organic porous media, gas can be stored as compressed  uid

    inside the pores or it can be adsorbed by the solid matrix. Similar to

    surface tension, adsorption is a consequence of surface energy

    (Gregg and Sing, 1982), which causes gas molecules to get bounded

    to the surface of the rock grains. The gas adsorption in shale-gas

    system is primarily controlled by the presence of organic matter

    and the gas adsorption capacity depends on TOC (Total Organic

    Carbon), organic matter type, thermal maturity and clay minerals

    (Ambrose et al., 2010; Chalmers and Bustin, 2008; Hill et al., 2007;

     Jarvie et al., 2007; Jenkins and Boyer, 2008; Passey et al., 2010; Ross

    and Bustin, 2008). Generally, the higher the TOC content, the

    greater the gas adsorption capacity (Zhang et al., 2013). In addition,

    a large number of nanopores lead to signicant nanoporosity in

    shale formations, which increases the gas adsorption surface area

    substantially ( Javadpour et al., 2007). The amount of adsorbed gas

    varies from 35 to 58% (Barnett Shale, USA) up to 60e85% (Lewis

    Shale, USA) of total gas initial in-place ( Darishchev et al., 2013).

    Presently, the only method for accurately determining the

    adsorbed gas in a formation is through core sampling and analysis.

    However, understanding the effects that initial adsorption, and

    moreover, desorption has on gas production will increase the

    effectiveness of reservoir management in these challenging

    environments.In shale formations, the total gas in-place (GIP) is determined by

    free gas in-place and adsorbed gas inplace. Adsorbed gas can be the

    dominant in-place resource in highly organically rich shales. Free

    gas becomes the dominant in-place resource in high clastic content

    shales. The calculation of free gas in-place for a given areal extent is

    governed, to a large extent, by pressure, temperature, gas-lled

    porosity and net organically-rich shale thickness. A Langmuir

    isotherm (1916) can be established for a prospective area of shale

    basin using available data on TOC and on thermal maturity to

    establish the Langmuir volume (V L) and the Langmuir pressure (P L).

    Adsorbed gas in-place is then calculated using the formula below:

    mad  ¼ r

    mr

     g 0V L

    P  þ P L (1)

    where mad is the gas adsorption mass per unit volume of formation

    (kg m3),  rm is shale matrix density (kg m3),  r g 0  is gas density at

    standard condition (kg/scf). In general, Langmuir volume  V L  (scf/

    ton) is a function of the organic richness and thermal maturity of 

    the shale and the Langmuir pressure   P L(Pa) is a function of how

    readily the gas can be adsorbed on the organics in the shale matrix

    and how readily it is released as a function of a  nite decrease in

    pressure.

    The Langmuir isotherm (Fig. 1) gives us an idea of how the

    adsorbed gas, free gas and total gas capacity of the shale reservoir

    relates with the pressure in terms of gas content. We can also

    deduce that the reservoir pressure must be suf ciently low to

    liberate the adsorbed gas and the ultimate recoverable amount of 

    gas is largely a function of the adsorbed gas that can be released

    (desorbed) in organically rich shales. Because most adsorbed gas

    can only be released at low reservoir pressure, due to the extremely

    low permeability in shale matrix, even with hydraulic fracturing, it

    would take considerable production time for the average pressure

    within the drainage area to drop to a level where most of the

    adsorbed gas can be liberated, and the production rate may have

    already reached the economical shut-in limits by then.

    Thermal stimulation techniques that can increase the formation

    temperature around the hydraulic fracture during production can

    be utilized as a potential method to enhance ultimate recovery

    from shale gas reservoir by altering shale gas desorption behavior.

    This will enhance desorption rate and recovery of a larger amount

    of gas prior to reaching reservoir pressure economic limits.

    Numerous authors have investigated how to enhance hydro-

    carbon recovery through thermal treatment, such as in-situ com-

    bustion (Hascakir et al., 2013), using electromagnetic materials

    (Yahya et al., 2012), injecting uids including CO2 or steam through

    created hydraulic fractures (Biglarbigi et al., 2007; Thoram et al.,

    2011).

    Symington et al. (2006) proposed an electric heating method for

    converting oil shale to producible oil and gas through heating theoil shale in situ by hydraulically fracturing the oil shale and  lling

    the fracture with electrically conductive material to form a heating

    element. The temperature in the hydraulic fracture could be stim-

    ulated up to a temperature of 400   C (752   F).

    Another possible way to heat the hydraulic fracture directly is to

    use nanoparticles forincreased coal seam gas, shale gasand oilshale

    production via enhanced electromagnetic heating of hydrocarbon

    deposits. Already, microwave applications in oil sands bitumen and

    shale oil production and in petroleum upgrading are gaining

    considerable interest in recent years. Energy companies and petro-

    leum researchers have been working on a variety of unconventional

    technologies such as microwave and radio frequency (RF) energies

    torecoveroil and gas from oil shale (Mutyala et al., 2010). Underthe

    oil shale extraction scenario,auxiliary wellsare drilled intothe shalestrata usingstandardoil-industry equipment, and then RF antennas,

    or transmitters, are lowered into wells. The antennas then transmit

    RF energy to heat the buried shale rock. This results in the volatili-

    zation of water, which in turn, results in the micro-fracturing of the

    formation, enhancing product recovery uniformly. Microwaves can

    generate heat faster than convection heating and it is observed that

    shale can be adequately heated to extract oil within a month or two

    of beginning production activities, rather than the year or longer for

    other methods. Nanoparticles in the form of nanouids have been

    used for enhanced oil recovery applications and it has shown that

    due to absorption of electromagnetic waves by the cobalt ferrite

    nanoparticles that were used, oil viscosity was reduced resulting in

    an increase of oil recovery (Yahya et al., 2012). However, all of these

    studies only focus on improving heavy oil/shale oil recovery byincreasing formation temperature to convert kerogen into oil/gas

    andreduce oilviscosity. Few studies have exploredthe possibility of 

    enhancing shale gas recovery with similar techniques. Now the

    question arises: Can these thermal treatments be applied to shale

    gas reservoirs? If we can increase the temperature within the hy-

    draulicfracture with similar methods,how much will itaffectthegas

    desorption mechanism in shale matrix and the ultimate recoverable

    amount of gas?

    In this study, we proposed a method to evaluate shale gas

    adsorption/desorption capacity as a function of pressure and tem-

    perature through regression of a Bi-Langmuir model based on data

    from Langmuir isotherm. An unconventional, fully coupled, hy-

    draulic fractured shale gas reservoir simulator was developed to

    investigate how the fracture temperature will impact the shale gasFig. 1.   Combining free and adsorbed gas for total gas in-place (Kuuskraa et al., 2011).

    H. Wang et al. / Journal of Natural Gas Science and Engineering 21 (2014) 874e885   875

  • 8/9/2019 1-s2.0-S187551001400314X-main

    3/12

    production and what main factors can determine whether increase

    formation temperature can be successful. The volume of the

    reservoir that realizes an effect of the thermal stimulation will be

    referred to herein as the Stimulated Reservoir Volume (SRV) for the

    purposes of this paper. The results of this study can provide a

    foundation for future research efforts on cost-effective thermal

    stimulation techniques to enhance the ultimate shale gas recovery.

    2. Thermodynamic casting of shale gas adsorption/

    desorption

    Traditionally, Langmuir isotherms are widely used to model

    adsorption of coal bed methane and shale gas at a constant tem-

    perature, by measuring crushed rock samples in laboratory to

    describe the amount of adsorbed gas as a function of its pressure at

    constant temperature (Vasilache, 2010). This model is a simplied

    model of gas/solid adsorption which is based on four assumptions

    (Czepirski et al., 2000):

    1. All of the adsorption sites are equivalent and each site can only

    accommodate one molecule.

    2. The surface is energetically homogeneous and adsorbed mole-cules do not interact.

    3. There are no phase transitions.

    4. At the maximum adsorption, only a monolayer is formed.

    Adsorption only occurs on localized sites on the surface, not

    with other adsorbates.

    With these assumptions, the Langmuir isotherm equation can

    be rigorously derived from the theory of kinetics and thermody-

    namics, and is commonly written as (Adamson, 1982; Ruthven,

    1984):

    mad  ¼  rmr g 0V LK ðT ÞP 

    1 þ K ðT ÞP   (2)

    K (T ) is the adsorption equilibrium constant (1/Pa), which can be

    considered as temperature dependent:

    K ðT Þ ¼  k0T 12e

     E RT  (3)

    where k0 is a pre-exponential constant independent of temperature

    (1/Pa), T  is temperature in Kelvin, E  is the characteristic adsorption

    energy (J/mol) and R  is the universal gas constant (8.3145 J/mol K).

    However, regardless that this Langmuir model can describe

    monolayer gas adsorption on a homogeneous adsorbent well

    enough, it may not be suitable for gas/shale systems because

    different materials may contribute to gas adsorption.   Lu et al.

    (1995) investigated temperature dependent adsorption curves on

    shale samples and proposed a Bi-Langmuir model that can account

    for gas adsorption on both clay minerals and kerogen:

    mad  ¼  rmr g 0V L

     f 1

    K 1ðT ÞP 

    1 þ K 1ðT ÞP  þ f 2

    K 2ðT ÞP 

    1 þ K 2ðT ÞP 

      (4 e1)

    where   f i   is dened as the ratio of the amount of the   ith type of 

    adsorption at monolayer coverage to the total amount adsorbed at

    monolayer coverage and each adsorption site is assumed to follow

    the Langmuir equation. For two types of adsorption sites we have:

     f 1 þ f 2  ¼  1 (4 e2)

    K 1ðT Þ ¼  k1T 12e

    E 1RT  (4 e3)

    K 2ðT Þ ¼  k2T 12e

    E 2RT  (4 e4)

    The study indicated that the Bi-Langmuir model can   t the

    experimental data better than the Langmuir model when extrap-

    olating isotherms to desired temperatures. Even though the as-sumptions behind the Langmuir isotherm model may not always

    fully depict the real phenomenon, it is still the most common

    isotherm equation to use in industry due to its simplicity and its

    ability to  t a variety of adsorption data.

    In this study, we only investigate the thermal stimulation effects

    based on the Bi-Langmuir model obtained using regression from

    Langmuir isotherm data, we also assume the gas adsorption and

    desorption curves coincide without hysteresis. For a more complex

    adsorption model that can be applied to a specic case, which

    needs to consider the effects of multiple hydrocarbon components,

    multi-layer adsorption, nonhomogeneous adsorbents and phase

    change behavior, the reader should refer to various other ap-

    proaches that have been proposed by numerous authors (Ambrose

    et al., 2011; Das et al., 2012; Fathi and Akkutlu, 2009; Hartman et al.,

    2011; Ji et al., 2012; Leahy-Dios et al., 2011; Li et al., 2013; Zhang

    et al., 2013).

    The  ve unknown independent parameters  f 1,  k1,  k2,  E 1  and  E 2can be determined from a Langmuir isotherm curve at provided

    temperature condition by regression. Thus, the temperature effects

    can be included to describe shale gas adsorption capacity.  Fig. 2

    shows an example of temperature effects on gas adsorption ca-

    pacity: the original Langmuir isotherm is provided at the temper-

    ature of 150   F with Langmuir volume of 400 scf/ton and Langmuir

    pressure of 500 psi, and the adsorption capacity is extrapolated to

    the temperature of 400   F by  tting Eq. (4)   against Eq. (1). It can

    also be observed that with increasedtemperature, large amounts of 

    extra gas can be released even at high pressure.  Table 1 shows the

    estimation of parameters in the Bi-Langmuir model based on the

    provided Langmuir isotherm curve.

    3. Mathematical model

    The physical process of thermal treatment for shale gas reser-

    voirs involves uid ow within the formation matrix and hydraulic

    fracture, shale gas adsorption and desorption as a function of 

    pressure and temperature, real gas properties affected by pressure

    Fig. 2.  Gas adsorption capacity with different temperatures.

     Table 1

    Estimates of parameters in Bi-Langmuir model.

     f 1   k1 (1/psi)   E 1 (kcal/mol)   k2 (1/psi)   E 2  (kcal/mol)

    0.4017 0.0014   2.2032 0.000022262   5.0015

    H. Wang et al. / Journal of Natural Gas Science and Engineering 21 (2014) 874e885876

  • 8/9/2019 1-s2.0-S187551001400314X-main

    4/12

    and temperature, and heat transfer process between the hydraulic

    fracture and shale formation.

    Many shale gas reservoirs can have matrix permeability values

    in the range of tens to hundreds of nanodarcies. Due to the ultra-

    ne pore structure, non-Darcy  ow mechanisms such as Knudsen

    diffusion and/or gas slippage effects may affect matrix perme-

    ability, depending on a combination of pressure-temperature

    conditions, pore structure and gas properties. Even though a vari-

    ety of correlations and models have been proposed in recent years

    to account for the apparent permeability enhancement in nanopore

    structures (Florence et al., 2007; Javadpouret al., 2007; Civan, 2010;

    Michel et al., 2011; Fathi et al., 2012; Sakhaee-Pour and Bryant,

    2012), given that our main objectiveis to characterize the enhanced

    recovery in shale formations by heating through hydraulic frac-

    tures, this study will still be based on Darcy's Law to determine gas

    ow rate in shale matrix.

    q g  ¼ k  g m g 

    $VP    (5)

    where q g is velocity vector of gas phase (m/s),k  g is gas permeability

    vector (md), and  m g   is gas viscosity (Pa s).The continuity equation in shale formation is:

    vm

    vt   þ V$

    r g q g 

     ¼  Q m   (6)

    where m is total gas content per unit volume (kg m3), r g is local gas

    density (kg m3), Q m is the source term (kg m3 s1) and t   is time

    (s).

    The total gas content m   is a combination of two parts:

    m ¼  r g ∅m þ mad   (7)

    r g Øm is free gas mass in shale pore space (kg m3), and adsorption

    gas   mad

      can be determined by Eq.   (4)  at given temperature and

    pressure.

    In situ gas density can be calculated by the real gas law:

    r g  ¼  PM 

     ZRT   (8)

    where M  is average molecular weight of mixed gas (kg/mol) and Z-

    factor can be calculated with pseudo reduced pressure ( p pr ) and

    pseudo reduced temperature (T  pr ) of mixed gas by the correlation

    (Mahmoud, 2013):

     Z  ¼ð0:702e2:5T  pr Þð p2 pr Þ ð5:524e2:5T  pr Þð p pr Þ

    þ ð0:044T 2 pr   0:164T  pr  þ 1:15Þ(9)

    Among all gas properties, gas viscosity, which is usually char-

    acterized by some available correlations, still remains uncertain.

    Values of gas viscosities are used to model gas mobility in reser-

    voirs and have a signicant impact on reserves estimation during

    eld development planning. 1% error in gas viscosity can lead to 1%

    error in gas   ow rate prediction, which in a large scale of gas

    production may severely underestimate or overestimate recovery

    (Davani et al., 2009).   Lee et al. (1966)  proposed a correlation to

    calculate gas viscosity at temperatures from 100 to 340   F and

    pressure from 100 to 8000 psi.  Viswanathan (2007) measured the

    viscosity of pure methane at pressure from 5000 to 30,000 psi and

    temperatures from 100 to 400   F, and modied the correlation by

    Lee et al. (1966). Based on measurements and results of 

    Viswanathan (2007):

    m g  ¼ 104Ke X r

    Y  g  (10 e1)

    where

    K  ¼ ð5:0512 0:2888M ÞT 1:832

    443:8 þ 12:9M  þ T   (10 e2)

     X  ¼ 6:1166 þ

    3084:9437

     þ 0:3938M    (10 e3)

    Y  ¼  0:5893 þ 0:1563 X    (10 e4)

    Tangential derivatives (Chen et al., 2013) are used to dene the

    ow along the interior boundary representing a hydraulic fracture

    within the porous medium. Since a hydraulic fracture itself is  lled

    with proppant and the gas   ow rate is relatively slow due to

    extremely low formation permeability,   ow behavior inside the

    hydraulic fracture can be described by Darcy's Law:

    q f   ¼ k  f m g 

    $d f VT P    (11)

    where   q f   is the volume   ow rate vector per unit length in the

    fracture (m3/s m),  k f   is fracture permeability vector (md),   VTP   is

    pressure gradient tangent to fracture surface.

    The continuity equation along the fracture reects the material

    balance:

    d f v∅ f r g 

    vt   þ VT $

    r g q f 

     ¼  d f Q  f    (12)

    where d f  is hydraulic fracture width, Øf  is the fracture porosity, and

    Q  f  is the mass source term (kg m3 s1), which can be calculated by

    adding mass ow rate per unit volume from two fracture walls (left

    and right) by:

    Q  f   ¼ Q  f lef t 

     þ  Q  f right

      (13 e1)

    Q  f left

     ¼ k  g u g 

    vP leftvnleft

    (13 e2)

    Q  f right

     ¼ k  g u g 

    vP rightvnright

    (13 e3)

    where  n  is the vector perpendicular to fracture surface.

    When the hydraulic fracture is directly heated, the fracture

    surface is considered as a constant temperature boundary condi-

    tion. Heat transfer between the hydraulic fracture and formation

    matrix is governed by a thermal diffusion equation in an isotropic

    porous medium; radiative effects, viscous dissipation, and the work

    done by pressure changes are negligible. Considering an elemental

    volume of a porous medium we have, for the matrix:

    ð1 ∅mÞrmC mvT mvt 

      þ ð1 ∅mÞV$ðk mVT mÞ ¼ ð1 ∅mÞQ m

    (14)

    and for the gas phase:

    ∅mr g C  p; g vT  g vt 

      þ∅mV$

    k  g VT  g 

     þ  r g C  p; g q g $VT  g  ¼ ∅mQ  g    (15)

    where  r  is the density (kg m3), C m is the heat capacity of the rock

    matrix, C  p, g  is heat capacity at constant pressure of the gas phase (J/

    H. Wang et al. / Journal of Natural Gas Science and Engineering 21 (2014) 874e885   877

  • 8/9/2019 1-s2.0-S187551001400314X-main

    5/12

  • 8/9/2019 1-s2.0-S187551001400314X-main

    6/12

    temperature is imposed along the hydraulic fracture. The combi-

    nationof input parameters is designed to reecta relatively high gas

    adsorption level,so that nearly80% of initial gasin-place comesfrom

    adsorbed gas, this can be the case in organic rich shale formations

    such as Devonian shales (Schettler and Parmely, 1991).

    5. Results and analysis

    In this section, we present results of the analysis to demonstrate

    howheating a hydraulic fracture canimpact the release of adsorbed

    gas in a given SRV and what factors can inuence long term pro-

    duction performance. Each SRV has its own fracture, each with an

    equal geometry of half-length and permeability.

    Fig. 4(a) shows the temperature, pressure, viscosity and density

    proles in the simulated SRV, all of which are time-space depen-

    dent during production. After 1 year's production, the front edge of 

    temperature diffusion has reached the virtual SRV boundary while

    the pressure propagation has not touched the boundary and the

    ow regime still exhibits transient linear   ow. Gas viscosity and

    density values are lowest in the high temperature and low pressure

    zone around the hydraulic fracture.

    Fig. 4(b)   shows the same proles without thermal stimula-

    tion. Comparing these results with the ones in  Fig. 4(a), it can

    be observed that the gas density is higher and the viscosity is

    slightly lower in the pressure depleted zone due to lower

    formation temperature. More important, it can be also noticed

    that the pressure propagation has reached the virtual SRV 

    boundary and the fracture started to deplete its SRV under

    pseudo-steady state after 1 year's production. This indicates

    that the extra adsorbed gas released within the thermally

    stimulated zone provides a self-induced pressure maintenance

    effect.

    Fig. 4.  (a). Temperature-pressure-viscosity-density distribution after 1 year of production with thermal stimulation. (b). Temperature-pressure-viscosity-density distribution after

    1 year of production without thermal stimulation.

    H. Wang et al. / Journal of Natural Gas Science and Engineering 21 (2014) 874e885   879

  • 8/9/2019 1-s2.0-S187551001400314X-main

    7/12

    Fig. 5 shows the percentage of adsorbed gas produced after 10

    years' production with and without thermal treatment. In the case

    without thermal stimulation, nearly 45 percent of original adsor-

    bed gas has been released from the vicinity near to the hydraulic

    fracture (within a distance of 15 ft). When the hydraulic fracture is

    heated up to 400   F, nearly 85 percent of original adsorbed gas can

    be released from this near fracture region.

    Fig. 6 shows total production rate from the simulated horizontalwell with an assumed count of 20 transverse hydraulic fractures,

    throughout therst ve years of production. It can be observed that

    even though the thermal stimulation production rate is higher than

    that without thermal stimulation, both rates decline rapidly during

    the  rst year of production. The thermal treatment does not pre-

    vent substantial reduction in production rate in early stages, as can

    be observed in both cases, where production rate decreased nearly

    80% over the  rst 2 years of production.

    5.1. Impact of stimulation temperature

    In order to investigate how the magnitude of stimulation tem-

    perature can impact the long term production, a constant tem-

    perature applied along the hydraulic fracture was modeled from200   F to 400   F with intervals of 50   F for a reservoir with initial

    temperature of 150   F.

    The simulation results are shown in Fig. 7 and it can be observed

    that by increasing the stimulation temperature along the hydraulic

    fracture, long term cumulative production can be enhanced. The

    larger the temperature difference between the stimulation tem-

    perature and initial reservoir temperature, larger is the volume of 

    desorbed gas that can be recovered. With 400   F of stimulation

    temperature, more than 20% additional gas can be recovered after

    20 years of production.

    5.2. Impact of shale thermal properties

    The formation thermal properties such as heat capacity and heat

    conductivity used in Eq. (17) may impact the ef ciency of thermal

    stimulation. These properties are a function of several parameters

    of a particular shale formation, including composition, porosity,

    water content, clay content and even changes with temperature

    and pressure.   Gilliam and Morgan (1987)  conducted laboratory

    experiments and a systematic study on shale samples form Devo-

    nian shale, Pierre shale and Green River shale formation and

    Railsback (2011)   compiled data of thermal conductivity of sedi-

    mentary rocks as shown in  Fig. 8.Based on their work and the ranges of temperature and

    pressure investigated in this study, the heat capacity in our

    model was set from 500 J/(kg K) to 1500 J/(kg K) and heat con-

    ductivity was set from 1 W/(m K) to 6 W/(m K), to reect the

    possible ranges of these thermal properties in shale and the in-

    uence of laminated interbeded formations or layers of sands

    and carbonate.

    Fig. 9   shows that the formation heat capacity has limited

    impact in the early stages of production and has negligible im-

    pacts on thermal stimulation effects after 20 years of production,

    because the whole SRV region will be heated up to a desired

    temperature regardless of formation heat capacity if given enough

    time for the heat diffusion process to occur. It should be

    mentioned however, that even though the formation heat ca-pacity will not make much difference in the cumulative produc-

    tion while the fracture temperature is kept constant, it has a huge

    impact on how much energy is required to heat up the formation.

    The smaller the heat capacity, the less input energy is required for

    heating the formation.

    Fig. 10 demonstrates that heat conductivity impacts the cumu-

    lative production, especially in the 2e12 years window. The smaller

    the heat conductivity, the faster the SRV region can be heated up to

    the desired temperature. After the whole SRV area reaches the

    stimulation temperature, the heat conductivity becomes less

    important.

    Fig. 5.  Percentage of adsorbed gas produced after 10 years of production.

    Fig. 6.  Production rate decline during 5 years of production.   Fig. 7.  Cumulative production with different stimulation Temperature  T  f .

    H. Wang et al. / Journal of Natural Gas Science and Engineering 21 (2014) 874e885880

  • 8/9/2019 1-s2.0-S187551001400314X-main

    8/12

    5.3. Impact of fracture spacing 

    The spacing between perforations and the number and orien-

    tation of transverse fractures have a major impact on well pro-

    duction, and the fracture spacing should avoid stress interference

    between transverse fractures in order to obtain optimal fracture

    geometry (Roussel and Sharma, 2011; Morrill and Miskimins, 2012;

    Wang, H et al., 2014). Hydraulic fracture spacing is a critical

    parameter in the design of multi-stage transverse fractures in

    horizontal wells. Here we investigate how fracture spacing will

    impact the thermal stimulation, starting from our previously

    studied base case, i.e. a horizontal well with 20 transverse hydraulic

    fractures with 120 ft fracturing spacing.

    Fig. 11 shows the cumulative production under different fracturespacing, keeping constant the total reservoir drainage area, with

    and without thermal stimulation. It is not surprising that the

    smaller the fracture spacing (i.e., the larger number of transverse

    hydraulic fractures in the drainage area), and the larger the thermal

    stimulation input, the more gas is recovered. Appropriate fracture

    spacing should therefore be designed based on the impact of geo-

    mechanical effects (stress interference effects) and economic ef-

    fects (Net Present Value of heating vs. production).

    5.4. Impact of formation permeability

    Permeability is one of the most fundamental properties of any

    reservoir rock required for modeling hydrocarbon production and

    the optimal hydraulic fracture design is also highly dependent on

    the reservoir permeability (Economides et al., 2002). Many shale

    gas reservoirs have extremely low matrix permeability values in

    the range of tens to hundreds of nanodarcies, and the low forma-

    tion permeability can adversely impact the pressure diffusion

    process so that the transient  ow in SRV may dominate the entireproduction life and the average pressure in the SRV will not be able

    to decline to a level that allows suf cient desorption before pro-

    duction rates reach economical limits.

    Fig. 12   shows the cumulative production under different for-

    mation permeability scenarios with and without thermal stimula-

    tion. It can be observed that the formation permeability seems to

    have a negligible impact on how much extra gas (around 2.8 BSCF

    for all cases) can be produced after 20 years of production, but the

    ratio of cumulative production under heating to cumulative pro-

    duction without heating decreases as formation permeability

    increases.

    5.5. Impact of bottom hole pressure

    In order to mimic possible surface production conditions and

    constraints on wellhead  owing pressure, the corresponding bot-

    tom hole pressure (BHP), which is a function of the hydrostatic

    pressure difference and the frictional pressure losses in the tubing

    string, is also subject to constraints.

    Fig. 13   shows the effects of heating shale formation with

    different values of BHP. It can be observed that after 20 years pro-

    duction, 50% extra gas can be produced when the BHP is 2000 psi

    while 20% extra gas can be produced when the BHP is 500 psi. The

    is because with higher BHP, the average pressure in the SRV will

    never decline to a level that most adsorbed gas can be liberated, so

    the temperature effects play a more important role in releasing

    adsorbed gas at high BHP.

    Fig. 8.  Thermal conductivity of different sedimentary rocks (modied from Railsback,

    2011).

    Fig. 9.  Cumulative production with different formation heat capacity C 

    m.

    Fig. 10.  Cumulative production with different formation heat conductivity  k m.

    Fig. 11.  Cumulative production with different fracture spacing  Y e.

    H. Wang et al. / Journal of Natural Gas Science and Engineering 21 (2014) 874e885   881

  • 8/9/2019 1-s2.0-S187551001400314X-main

    9/12

    5.6. Impact of Langmuir volume

    According to Eq.   (4), Langmuir volume,   V L, determines how

    much gas can be ultimately adsorbed by the rock matrix, which

    reects shale formation's intrinsic adsorption capacity. Altering the

    value of Langmuir volume is equivalent to altering the amount of 

    adsorbed gas initially in place, but the value of Langmuir volume

    will not impact the shape of the adsorption curve and the results of 

    the ve regression parameters of Bi-Langmuir model, as long as the

    Langmuir pressure,  P L, is  xed. Fig. 14  shows the cumulative pro-

    duction with different Langmuir volume. The results indicate that

    there is considerable improvement in recovery (~30% at 400 scf/

    ton) with higher Langmuir volumes, however with smaller

    Langmuir volume, less adsorption gas is initially in place and less

    extra recovery is achieved by heating the formation. When the

    Langmuir volume decreases to 100 scf/ton, the additional gas

    recovered by thermal stimulation is negligible.

    5.7. Impact of Langmuir pressure

    As discussed in the previous section, the Langmuir pressure,  P L,

    is a function of how readily the adsorbed gas on the organics in the

    shale matrix is released as a function of a  nite decrease in pres-sure. Different values of Langmuir pressure lead to different shapes

    of gas adsorption curves. Table 3 shows the results of the regression

    performed to determine the parameters in the Bi-Langmuir model

    with different Langmuir pressures.

    The larger the Langmuir pressure, the smoother the adsorption

    curve will look, as shown in Fig.15. It can also be noticed that with

    smaller Langmuir pressure, the pressure has to be suf ciently low

    in order to release the majority of adsorption gas.

    Fig. 16 demonstrates that in general, larger amounts of gas can

    be produced with higher Langmuir pressure, due to the fact that

    more gas will be desorbed even at relatively higher pressure. It can

    also be observed that with higher Langmuir pressure, less gas will

    be desorbed by effects of the thermal stimulation; this is because if 

    Fig.12.  Cumulative production with different formation permeability  k g ,  P L ¼ 500 psi.

    Fig. 13.  Cumulative production with different bottom hole pressure.

    Fig. 14.  Cumulative production with different Langmuir volume V  L .

     Table 3

    Estimates of parameters in Bi-Langmuir model with different Langmuir pressures.

    Langmuir

    pressure

     f 1   k1 (1/psi)   E 1 (kcal/

    mol)

    k2 (1/psi)   E 2 (kcal/

    mol)

    500 psi 0.4017 0.0014   2.2032 0.000022262   5.0015

    1000 psi 0.4014 0.000941   2.0178 0.0000108   5.0004

    2000 psi 0.2673 0.000695   1.7723 0.0000079036 4.7423

    Fig. 15.  Gas adsorption curve with different Langmuir pressures  P L.

    Fig. 16.  Cumulative production with different Langmuir pressures  P L .

    H. Wang et al. / Journal of Natural Gas Science and Engineering 21 (2014) 874e885882

  • 8/9/2019 1-s2.0-S187551001400314X-main

    10/12

    most gas can be released at a higher reservoir pressure, then the

    stimulation temperature will become less important.

    Fig. 17 shows the cumulative production with different forma-

    tion permeability when Langmuir pressure is 2000 psi. When

    compared to Fig. 12 where the Langmuir pressure is 500 psi, for-

    mation permeability plays a more important role in the effective-

    ness of thermal stimulation, the amount of extra gas recovered by

    thermal stimulation decreases with increasing formation perme-

    ability. This is because when the Langmuir pressure is high, more

    adsorbed gas can be released at a higher pressure and the higher

    formation permeability can promote the pressure diffusion process

    which leads to more desorption due to pressure depletion rather

    than temperature elevation.

    6. Discussions

    The fundamental rationale behind thermal stimulation of shale

    gas reservoirs is to promote the desorption of the adsorbed gas by

    increasing the formation temperature which alters gas desorption

    behavior.

    The most important parameters for the effectiveness of thermal

    stimulation are Langmuir volume, Langmuir pressure, stimulationtemperature, fracture spacing, reservoir permeability and bottom

    hole pressure. Langmuir volume and Langmuir pressure determine

    the initial adsorbed gas in place and the shape of the desorption

    curve. An ideal candidate for thermal stimulation should have large

    Langmuir volume and relatively low Langmuir pressure. In this

    study, stimulation temperature in the hydraulic fracture has been

    modeled up to 400  F due to limited correlations for gas viscosity at

    high temperature. Nevertheless, it was found that the higher the

    stimulation temperature, the larger the ultimate recoverable vol-

    ume of gas.

    Fracture spacing determines how far the virtual boundary will

    be away from the hydraulic fracture, the closer the fracture spacing

    is, the faster the SRV can be heated upto a desired temperature. The

    formation permeability determines how long the transient   owregime in the SRV will last and how fast the average pressure will

    decline. Its impact on thermal stimulation results depends on

    Langmuir pressure as the higher the Langmuir pressure, the deeper

    the inuence of formation permeability. Based on the gas adsorp-

    tion curve, the value of the bottom hole pressure can determine the

    pressure drop prole in the drainage area and affect how much

    adsorbed gas can be ultimately produced due to pressure decline.

    To a lesser degree, thermal properties of shale formations inuence

    the production of thermally stimulated wells. Within the range of 

    possible values investigated in this study, the shale formation heat

    capacity has negligible impacts on the production and heat con-

    ductivity only has noticeable inuence on the effectiveness of 

    thermal stimulation at early years of production.

    Overall, the effectiveness of thermal stimulation is largely

    dependent on the volume of adsorbed gas that can be recovered by

    pressure depletion alone. If the average pressure in the SRV can

    drop to a suf ciently low level to allow gas desorption within a

    reasonable time frame, then the rationale behind thermal stimu-

    lation will primarily be to increase gas production rate to maximize

    NPV of the assets. If the average pressure in the SRV cannot reach a

    suf ciently low level to allow desorption, then thermal stimulation

    has great potential to be used as an effective recovery method to

    enhance ultimate recovery in shale reservoirs.

    Future research efforts should target laboratory experiment to

    explore pressure and temperature dependency, mixed gas

    adsorption behavior in an isotropic formations, with laminated and

    variable lithologyand the feasibility of using electrically conductive

    material or nano-particle coated proppant to heat the shale for-

    mation in-situ along created hydraulic fractures via electric or

    electromagnetic heating. Additionally, shear induced fracture slip

    and formation permeability alteration due to thermal stress should

    also be investigated. It should be emphasized that this paper is not

    intended to discuss the techniques of heating hydraulic fracture indetail or to determine when to start thermal stimulation process or

    how long the stimulation process should last from an economical

    point of view. The purpose of this paper is to provide a feasibility

    study of enhanced shale gas recovery by heating along hydraulic

    fractures, and the results of this study can be used as general

    guidance for future research efforts on developing cost effective

    thermal recovery method in shale formation.

    7. Conclusions

    In this study, we proposed a method to evaluate shale gas

    desorption as a function of pressure and temperature, by regression

    of a Bi-Langmuir model based on Langmuir isotherm data. A fully

    coupled unconventional reservoir simulator was developed toinvestigate how thermal stimulation in hydraulically fractured

    formations can impact shale gas recovery and what factors most

    impact the effectiveness of thermal stimulation. The results of this

    study show that thermal stimulation can promote gas recovery and

    the production rate of the desorbed gas is primarily determined by

    the gas desorption curve, stimulation temperature, fracture

    spacing, reservoir permeability, and bottom hole pressure. A good

    candidate shale gas reservoir for thermal treatment should have

    large Langmuir volume, low Langmuir pressure, relatively small

    heat capacity, and high thermal conductivity. On the operational

    side, fracture spacing and stimulation temperature are the only two

    variables that we can control during treatment design and

    execution.

    Nomenclature

    C m   Matrix heat capacity at constant pressure, J/(kg K)

    d f    Fracture width, m

    E    Characteristic adsorption energy, J/mol

    H    Net pay thickness, m

    k0   Regression constant, 1/Pa

    k  f    Fracture permeability vector, md

    k  g    Gas permeability vector, md

    k m   Matrix thermal conductivity, W/(m K)

    K (T ) Adsorption equilibrium constant, 1/Pa

    m   Total gas content, kg m3

    mad   Gas adsorption mass per unit volume, kg m3

    M    Average molecular weight, kg/molFig. 17.  Cumulative production with formation permeability k 

    g , P  L   ¼

     2000 psi.

    H. Wang et al. / Journal of Natural Gas Science and Engineering 21 (2014) 874e885   883

  • 8/9/2019 1-s2.0-S187551001400314X-main

    11/12

    n   Normal vector

     p pr    Pseudo reduced pressure

    P    Reservoir Pressure, Pa

    P L   Langmuir pressure, Pa

    q f    ow rate vector in the fracture, m3/s m

    q g    Velocity vector of gas phase, m/s

    Q    Heat source term in matrix, J m3 s1

    Q  f    Mass source term in fracture, kg m3 s1

    Q m   Mass source term in matrix, kg m3 s1

    R   Universal gas constant, 8.3145 J/mol K

    r w   Well radius, m

    sc    Choke skin

    T    Reservoir temperature, K

    T  f    Thermal stimulation temperature along hydraulic

    fracture, K

    T  pr    Pseudo reduced temperature

    V L   Langmuir volume, scf/ton

     X e   SRV length parallel to the hydraulic fracture, m

    Y e   Fracturing spacing, m

    m g    Gas viscosity, Pa s

    r g    Gas density, kg m3

    r g 0   Gas density at standard condition, kg/scf 

    rm   Matrix density, kg m3

    Ø f    Fracture porosity

    Øm   Matrix porosity

    References

    Adamson, W.,1982. Physical Chemistry of Surfaces, fourth ed. John Wiley and Sons,New York City.

    Ambrose, R.J., Hartman, R.C., Diaz-Campos, M., 2010. New pore-scale considerationsfor shale gas in place calculations. In: SPE Unconventional Gas Conference,23e25 February, Pittsburgh, Pennsylvania, USA.

    Ambrose, R.J., Hartman, R.C., Akkutlu, I.Y., 2011. Multi-component sorbed-phaseconsiderations for shale gas-in-place calculations. In: Paper Presented at theSPE Production Operations Symposium Held in Oklahoma City, Oklahoma,March 27e29.

    Biglarbigi, K., Dammer, A., Cusimano, J., Mohan, H., 2007. Potential for oil shaledevelopment in the United States. In: Paper Presented at the SPE AnnualTechnical Conference and Exhibition, Anaheim, California, U.S.A. November11e14.

    Chalmers, G.R.L., Bustin, R.M., 2008. Lower cretaceous gas shales in northeasternBritish Columbia, part II: evaluation of regional potential gas resources. Bull.Can. Petrol. Geol. 56, 22e61.

    Chen, B., Song, E., Cheng, X., 2013. Springer Series in Geomechanics and Geo-engineering, pp. 149e154.

    Civan, F., 2010. Effective correlation of apparent gas permeability in tight porousmedia. Transp. Porous Media 82 (2), 375e384.

    Czepirski, L., Balys, M.R., Komorowska-Czepirska, E., 2000. Some generalization of Langmuir adsorption isotherm. Internet J. Chem. 3 (14).

    Darishchev, A., Rouvroy, P., Lemouzy, P., 2013. On simulation of   ow in tight andshale gas reservoirs. In: Paper was Presented at the SPE Unconventional GasConference and Exhibition, Held at Muscat, Oman, January 28e30.

    Das, M., Jonk, R., Schelble, R., 2012. Effect of multicomponent adsorption/desorptionbehavior on Gas-In-Place (GIP) and Estimated Ultimate Recovery (EUR) in shalegas systems. In: Paper Presented at the SPE Annual Technical Conference and

    Exhibition Held in San Antonio, Texas, USA, October, 8e

    10.Davani, E., Ling, K., Teodoriu, C., McCain Jr., W.D., Falcone, G., 2009. Inaccurate gas

    viscosity at HP/HT conditions and its effect on unconventional gas reservesestimation. In: Paper Presented at the SPE Latin American and Caribbean Pe-troleum Engineering Conference Held in Cartagena, Colombia, 31 Maye3 June.

    Economides, M.J., Oligney, R.E., Valko, P., 2002. Unied Fracture Design. Orsa Press.Fathi, E., Akkutlu, I.Y., 2009. Matrix heterogeneity effects on gas transport and

    adsorption in coalbed andshale gasreservoirs.Transp.Porous Med.80, 281e304.Fathi, E., Tinni, A., Akkutlu, I.Y., 2012. Shale gas correction to Klinkenberg slip theory.

    In: Paper SPE 154977 Presented at the SPE Americas Unconventional ResourcesConferences at Pittsburgh, Pennsylvania 5e7 June.

    Florence, F.A., Rushing, J.A., Newsham, K.E., Blasingame, T.A., 2007. Improvedpermeability prediction relations for low-permeability sands. In: Paper SPE107954 Presented at SPE Rocky Mountain Oil and Gas Technology SymposiumHeld at Denver, Colorado, April 16e18.

    Gilliam, T.M., Morgan, I.L., 1987. Shale: Measurement of Thermal Properties. U.SDepartment of Energy.

    Gregg, S.J., Sing, K.S.W., 1982. Adsorption, Surface Area and Porosity, second ed.

    Academic Press, New York. 303pp.

    Hartman, R.C., Ambrose, R.J., Akkutlu, I.Y., Clarkson, C.R., 2011. Shale gas-in-placecalculations part II-multicomponent gas adsorption effects. In: Paper Pre-sented at the North American Unconventional Gas Conference and ExhibitionHeld in the Woodlands, Texas, USA, June14e16.

    Hascakir, B., Ross, C., Castanier, L.M., Kovscek, A., 2013. Fuel formation and con-version during in-situ combustion of crude oil. SPE J. 18 (06), 1217 e1228.

    Hill, R.J., Zhang, E., Katz, B.J., Tang, Y., 2007. Modeling of gas generation from theBarnett Shale, Fort Worth Basin, Texas. AAPG Bull. 91, 501e521.

     Jarvie, D.M., Hill, R.J., Ruble, T.E., Pollastro, R.M., 2007. Unconventional shale-gassystems: the Mississippian Barnett Shale of north-central Texas as one model

    for thermogenic shale-gas assessment. AAPG Bull. 91, 475e

    499. Javadpour, F., Fisher, D., Unsworth, M., 2007. Nanoscale gas  ow in shale gas sedi-

    ments. J. Can. Petrol. Technol. 46 (10), 55e61. Jenkins, C., Boyer, C., 2008. Coalbed-and shale-gas reservoirs. J. Petrol. Technol. 60,

    92e99. Ji, L., Zhang, T., Milliken, K.L., Qu, J., Zhang, X., 2012. Experimental investigation of 

    main controls to methane adsorption in clay-rich rocks. Appl. Geochem. 27,2533e2545.

    Kuuskraa, V., Stevens, S., Leeuwen, T.V., Moodhe, K., 2011. World Shale Gas Report.U.S. Energy Information Administration.

    Lee, A.L., Gonzalez, M.H., Eakin, B.E., 1966. The viscosity of natural gases. JPT 997.Trans. AIME 237.

    Langmuir, I., 1916. The constitution and fundamental properties of solids and liq-uids. J. Am. Chem. Soc. 38 (11), 2221e2295.

    Leahy-Dios, A., Das, M., Agarwal, A., Kaminsky, R.D., 2011. Modeling of transportphenomena and multi-component sorption for shale gas and coal bed methanein an unstructured grid simulator. In: Paper Presented at the SPE AnnualTechnical Conference and Exhibition Held in Denver, Colorado, USA, 30October-2 November.

    Li, B., Mehmani, A., Chen, J., Georgi, D.T., Jin, G., 2013. The condition of capillarycondensation and its effects on adsorption isotherms of unconventional gascondensate reservoirs. In: Paper was Presented at the SPE Annual TechnicalConference and Exhibition Held in New Orleans, Louisiana, USA, 30 September-2 October.

    Lu, X.C., Li, F.C., Watson, A.T., 1995. Adsorption studies of natural gas storage inDevonian Shales. SPE Form. Eval. 10 (02), 109e113.

    Mahmoud, M.A., 2013. Development of a new correlation of gas compressibilityfactor (Z-factor) for high pressure gas reservoirs. In: SPE North Africa TechnicalConference and Exhibition Held in Cairo, Egypt, 15e17.

    Michel, G.G., Sigal, R.F., Civan, F., Devegowda, D., 2011. Parametric investigation of shale gas production considering nano-scale pore size distribution, formationfactor, and non-darcy ow mechanisms. In: Paper SPE 147438 Presented at SPEAnnual Technical Conference and Exhibition Held in Denver, Colorado, 30 Oct  e2 Nov.

    Morrill, J., Miskimins, J.L., 2012. Optimization of hydraulic fracture spacing in un-conventional shales. In: SPE Paper 152595 Presented at the SPE HydraulicFracturing Technology Conference Held in Woodlands, Texas, USA, 6e8

    February.Mukherjee, H., Economides, M.J., 1991. A Parametric Comparison of Horizontal andVertical Well Performance. SPE Paper 18303.

    Mutyala, S., Fairbridge, C., Pare, J.R.J., Belanger, J.M.R., Ng, S., Hawkins, R., 2010.Microwave applications to oil sands and petroleum: a review. Fuel ProcessTechnol. 91 (2), 127e135.

    Passey, Q., Bohacs, K., Esch, W., Klimentidis, R., Sinha, S., 2010. From oil-pronesource rock to gas-producing shale reservoir-geologic and petrophysical char-acterization of unconventional shale gas reservoirs. In: International Oil andGas Conference and Exhibition in China, 8e10 June. Society of Petroleum En-gineers, SPE, Beijing, China.

    Railsback, L.B., 2011. Petroleum Geoscience and Subsurface Geology.  http://www.gly.uga.edu/railsback/PGSG/PGSGmain.html.

    Ross, D.J.K., Bustin, R.M., 2008. Characterizing the shale gas resource potentialof Devonian-Mississippian strata in the Western Canada sedimentary ba-sin: application of an integrated formation evaluation. AAPG Bull. 92,87e12.

    Roussel, N.P., Sharma, M.M., 2011. Optimizing fracture spacing and sequencing inhorizontal well fracturing. SPE Prod. Oper. 26 (02), 87e12.

    Ruthven, D.M., 1984. Principles of Adsorption and Adsorption Processes. John Wileyand Sons, New York City.

    Sakhaee-Pour, A., Bryant, S.L., 2012. Gas permeability of shale. In: Paper SPE146944-PA, SPE Reservoir Evaluation and Engineering. August, pp. 401e409.

    Schettler, P.D., Parmely, C.R., 1991. Contributions to total storage capacity in Devo-nian Shales. In: Paper was Presented at the SPE Eastern Regional Meeting,Lexington, KY, October. 22e25.

    Symington, W.A., Olgaard, D.L., Otten, G.A., Phillips, T.C., Thomas, M.M., Yeakel, J.D.,2006. ExxonMobil's electrofrac process for in situ oil shale conversion. In: Paperwas Presented at the 26th Oil Shale Symposium, Held at the Colorado School of Mines in Golden, Colorado, October 16e18.

    Thoram, S., Ehlig-Economides, C.A., 2011. Heat transfer applications for the stimu-lated reservoir volume. In: Paper Presented at the SPE Annual Technical Con-ference and Exhibition, Denver, Colorado, USA. October 30eNovember 2.

    Vasilache, M.A., 2010. Fast and economic gas isotherm measurements using smallshale samples. In: Presentation at AAPG Annual Convention and Exhibition,New Orleans, Louisiana, April 11e14.

    Viswanathan, A., 2007. Viscosities of Natural Gases at High Pressures and HighTemperatures. MS thesis. Texas A and M University, College Station, Texas .

    H. Wang et al. / Journal of Natural Gas Science and Engineering 21 (2014) 874e885884

    http://refhub.elsevier.com/S1875-5100(14)00314-X/sref1http://refhub.elsevier.com/S1875-5100(14)00314-X/sref1http://refhub.elsevier.com/S1875-5100(14)00314-X/sref2http://refhub.elsevier.com/S1875-5100(14)00314-X/sref2http://refhub.elsevier.com/S1875-5100(14)00314-X/sref2http://refhub.elsevier.com/S1875-5100(14)00314-X/sref2http://refhub.elsevier.com/S1875-5100(14)00314-X/sref3http://refhub.elsevier.com/S1875-5100(14)00314-X/sref3http://refhub.elsevier.com/S1875-5100(14)00314-X/sref3http://refhub.elsevier.com/S1875-5100(14)00314-X/sref3http://refhub.elsevier.com/S1875-5100(14)00314-X/sref3http://refhub.elsevier.com/S1875-5100(14)00314-X/sref3http://refhub.elsevier.com/S1875-5100(14)00314-X/sref4http://refhub.elsevier.com/S1875-5100(14)00314-X/sref4http://refhub.elsevier.com/S1875-5100(14)00314-X/sref4http://refhub.elsevier.com/S1875-5100(14)00314-X/sref4http://refhub.elsevier.com/S1875-5100(14)00314-X/sref4http://refhub.elsevier.com/S1875-5100(14)00314-X/sref4http://refhub.elsevier.com/S1875-5100(14)00314-X/sref6http://refhub.elsevier.com/S1875-5100(14)00314-X/sref6http://refhub.elsevier.com/S1875-5100(14)00314-X/sref6http://refhub.elsevier.com/S1875-5100(14)00314-X/sref6http://refhub.elsevier.com/S1875-5100(14)00314-X/sref6http://refhub.elsevier.com/S1875-5100(14)00314-X/sref7http://refhub.elsevier.com/S1875-5100(14)00314-X/sref7http://refhub.elsevier.com/S1875-5100(14)00314-X/sref7http://refhub.elsevier.com/S1875-5100(14)00314-X/sref8http://refhub.elsevier.com/S1875-5100(14)00314-X/sref8http://refhub.elsevier.com/S1875-5100(14)00314-X/sref8http://refhub.elsevier.com/S1875-5100(14)00314-X/sref9http://refhub.elsevier.com/S1875-5100(14)00314-X/sref9http://refhub.elsevier.com/S1875-5100(14)00314-X/sref10http://refhub.elsevier.com/S1875-5100(14)00314-X/sref10http://refhub.elsevier.com/S1875-5100(14)00314-X/sref10http://refhub.elsevier.com/S1875-5100(14)00314-X/sref10http://refhub.elsevier.com/S1875-5100(14)00314-X/sref10http://refhub.elsevier.com/S1875-5100(14)00314-X/sref10http://refhub.elsevier.com/S1875-5100(14)00314-X/sref11http://refhub.elsevier.com/S1875-5100(14)00314-X/sref11http://refhub.elsevier.com/S1875-5100(14)00314-X/sref11http://refhub.elsevier.com/S1875-5100(14)00314-X/sref11http://refhub.elsevier.com/S1875-5100(14)00314-X/sref11http://refhub.elsevier.com/S1875-5100(14)00314-X/sref12http://refhub.elsevier.com/S1875-5100(14)00314-X/sref12http://refhub.elsevier.com/S1875-5100(14)00314-X/sref12http://refhub.elsevier.com/S1875-5100(14)00314-X/sref12http://refhub.elsevier.com/S1875-5100(14)00314-X/sref12http://refhub.elsevier.com/S1875-5100(14)00314-X/sref13http://refhub.elsevier.com/S1875-5100(14)00314-X/sref13http://refhub.elsevier.com/S1875-5100(14)00314-X/sref13http://refhub.elsevier.com/S1875-5100(14)00314-X/sref13http://refhub.elsevier.com/S1875-5100(14)00314-X/sref14http://refhub.elsevier.com/S1875-5100(14)00314-X/sref14http://refhub.elsevier.com/S1875-5100(14)00314-X/sref14http://refhub.elsevier.com/S1875-5100(14)00314-X/sref15http://refhub.elsevier.com/S1875-5100(14)00314-X/sref15http://refhub.elsevier.com/S1875-5100(14)00314-X/sref15http://refhub.elsevier.com/S1875-5100(14)00314-X/sref15http://refhub.elsevier.com/S1875-5100(14)00314-X/sref16http://refhub.elsevier.com/S1875-5100(14)00314-X/sref16http://refhub.elsevier.com/S1875-5100(14)00314-X/sref16http://refhub.elsevier.com/S1875-5100(14)00314-X/sref16http://refhub.elsevier.com/S1875-5100(14)00314-X/sref16http://refhub.elsevier.com/S1875-5100(14)00314-X/sref16http://refhub.elsevier.com/S1875-5100(14)00314-X/sref17http://refhub.elsevier.com/S1875-5100(14)00314-X/sref17http://refhub.elsevier.com/S1875-5100(14)00314-X/sref18http://refhub.elsevier.com/S1875-5100(14)00314-X/sref18http://refhub.elsevier.com/S1875-5100(14)00314-X/sref19http://refhub.elsevier.com/S1875-5100(14)00314-X/sref19http://refhub.elsevier.com/S1875-5100(14)00314-X/sref19http://refhub.elsevier.com/S1875-5100(14)00314-X/sref19http://refhub.elsevier.com/S1875-5100(14)00314-X/sref19http://refhub.elsevier.com/S1875-5100(14)00314-X/sref19http://refhub.elsevier.com/S1875-5100(14)00314-X/sref20http://refhub.elsevier.com/S1875-5100(14)00314-X/sref20http://refhub.elsevier.com/S1875-5100(14)00314-X/sref20http://refhub.elsevier.com/S1875-5100(14)00314-X/sref20http://refhub.elsevier.com/S1875-5100(14)00314-X/sref21http://refhub.elsevier.com/S1875-5100(14)00314-X/sref21http://refhub.elsevier.com/S1875-5100(14)00314-X/sref21http://refhub.elsevier.com/S1875-5100(14)00314-X/sref21http://refhub.elsevier.com/S1875-5100(14)00314-X/sref22http://refhub.elsevier.com/S1875-5100(14)00314-X/sref22http://refhub.elsevier.com/S1875-5100(14)00314-X/sref22http://refhub.elsevier.com/S1875-5100(14)00314-X/sref22http://refhub.elsevier.com/S1875-5100(14)00314-X/sref23http://refhub.elsevier.com/S1875-5100(14)00314-X/sref23http://refhub.elsevier.com/S1875-5100(14)00314-X/sref23http://refhub.elsevier.com/S1875-5100(14)00314-X/sref23http://refhub.elsevier.com/S1875-5100(14)00314-X/sref23http://refhub.elsevier.com/S1875-5100(14)00314-X/sref23http://refhub.elsevier.com/S1875-5100(14)00314-X/sref24http://refhub.elsevier.com/S1875-5100(14)00314-X/sref24http://refhub.elsevier.com/S1875-5100(14)00314-X/sref24http://refhub.elsevier.com/S1875-5100(14)00314-X/sref25http://refhub.elsevier.com/S1875-5100(14)00314-X/sref25http://refhub.elsevier.com/S1875-5100(14)00314-X/sref25http://refhub.elsevier.com/S1875-5100(14)00314-X/sref25http://refhub.elsevier.com/S1875-5100(14)00314-X/sref25http://refhub.elsevier.com/S1875-5100(14)00314-X/sref26http://refhub.elsevier.com/S1875-5100(14)00314-X/sref26http://refhub.elsevier.com/S1875-5100(14)00314-X/sref26http://refhub.elsevier.com/S1875-5100(14)00314-X/sref27http://refhub.elsevier.com/S1875-5100(14)00314-X/sref27http://refhub.elsevier.com/S1875-5100(14)00314-X/sref27http://refhub.elsevier.com/S1875-5100(14)00314-X/sref28http://refhub.elsevier.com/S1875-5100(14)00314-X/sref28http://refhub.elsevier.com/S1875-5100(14)00314-X/sref28http://refhub.elsevier.com/S1875-5100(14)00314-X/sref29http://refhub.elsevier.com/S1875-5100(14)00314-X/sref29http://refhub.elsevier.com/S1875-5100(14)00314-X/sref29http://refhub.elsevier.com/S1875-5100(14)00314-X/sref29http://refhub.elsevier.com/S1875-5100(14)00314-X/sref29http://refhub.elsevier.com/S1875-5100(14)00314-X/sref29http://refhub.elsevier.com/S1875-5100(14)00314-X/sref30http://refhub.elsevier.com/S1875-5100(14)00314-X/sref30http://refhub.elsevier.com/S1875-5100(14)00314-X/sref30http://refhub.elsevier.com/S1875-5100(14)00314-X/sref30http://refhub.elsevier.com/S1875-5100(14)00314-X/sref30http://refhub.elsevier.com/S1875-5100(14)00314-X/sref30http://refhub.elsevier.com/S1875-5100(14)00314-X/sref31http://refhub.elsevier.com/S1875-5100(14)00314-X/sref31http://refhub.elsevier.com/S1875-5100(14)00314-X/sref31http://refhub.elsevier.com/S1875-5100(14)00314-X/sref32http://refhub.elsevier.com/S1875-5100(14)00314-X/sref32http://refhub.elsevier.com/S1875-5100(14)00314-X/sref32http://refhub.elsevier.com/S1875-5100(14)00314-X/sref32http://refhub.elsevier.com/S1875-5100(14)00314-X/sref32http://refhub.elsevier.com/S1875-5100(14)00314-X/sref33http://refhub.elsevier.com/S1875-5100(14)00314-X/sref33http://refhub.elsevier.com/S1875-5100(14)00314-X/sref33http://refhub.elsevier.com/S1875-5100(14)00314-X/sref33http://refhub.elsevier.com/S1875-5100(14)00314-X/sref33http://refhub.elsevier.com/S1875-5100(14)00314-X/sref33http://refhub.elsevier.com/S1875-5100(14)00314-X/sref33http://refhub.elsevier.com/S1875-5100(14)00314-X/sref34http://refhub.elsevier.com/S1875-5100(14)00314-X/sref34http://refhub.elsevier.com/S1875-5100(14)00314-X/sref34http://refhub.elsevier.com/S1875-5100(14)00314-X/sref34http://refhub.elsevier.com/S1875-5100(14)00314-X/sref34http://refhub.elsevier.com/S1875-5100(14)00314-X/sref35http://refhub.elsevier.com/S1875-5100(14)00314-X/sref35http://refhub.elsevier.com/S1875-5100(14)00314-X/sref36http://refhub.elsevier.com/S1875-5100(14)00314-X/sref36http://refhub.elsevier.com/S1875-5100(14)00314-X/sref36http://refhub.elsevier.com/S1875-5100(14)00314-X/sref36http://refhub.elsevier.com/S1875-5100(14)00314-X/sref36http://refhub.elsevier.com/S1875-5100(14)00314-X/sref37http://refhub.elsevier.com/S1875-5100(14)00314-X/sref37http://refhub.elsevier.com/S1875-5100(14)00314-X/sref37http://refhub.elsevier.com/S1875-5100(14)00314-X/sref37http://refhub.elsevier.com/S1875-5100(14)00314-X/sref37http://refhub.elsevier.com/S1875-5100(14)00314-X/sref37http://www.gly.uga.edu/railsback/PGSG/PGSGmain.htmlhttp://www.gly.uga.edu/railsback/PGSG/PGSGmain.htmlhttp://refhub.elsevier.com/S1875-5100(14)00314-X/sref39http://refhub.elsevier.com/S1875-5100(14)00314-X/sref39http://refhub.elsevier.com/S1875-5100(14)00314-X/sref39http://refhub.elsevier.com/S1875-5100(14)00314-X/sref39http://refhub.elsevier.com/S1875-5100(14)00314-X/sref39http://refhub.elsevier.com/S1875-5100(14)00314-X/sref40http://refhub.elsevier.com/S1875-5100(14)00314-X/sref40http://refhub.elsevier.com/S1875-5100(14)00314-X/sref40http://refhub.elsevier.com/S1875-5100(14)00314-X/sref41http://refhub.elsevier.com/S1875-5100(14)00314-X/sref41http://refhub.elsevier.com/S1875-5100(14)00314-X/sref42http://refhub.elsevier.com/S1875-5100(14)00314-X/sref42http://refhub.elsevier.com/S1875-5100(14)00314-X/sref42http://refhub.elsevier.com/S1875-5100(14)00314-X/sref43http://refhub.elsevier.com/S1875-5100(14)00314-X/sref43http://refhub.elsevier.com/S1875-5100(14)00314-X/sref43http://refhub.elsevier.com/S1875-5100(14)00314-X/sref43http://refhub.elsevier.com/S1875-5100(14)00314-X/sref44http://refhub.elsevier.com/S1875-5100(14)00314-X/sref44http://refhub.elsevier.com/S1875-5100(14)00314-X/sref44http://refhub.elsevier.com/S1875-5100(14)00314-X/sref44http://refhub.elsevier.com/S1875-5100(14)00314-X/sref44http://refhub.elsevier.com/S1875-5100(14)00314-X/sref45http://refhub.elsevier.com/S1875-5100(14)00314-X/sref45http://refhub.elsevier.com/S1875-5100(14)00314-X/sref45http://refhub.elsevier.com/S1875-5100(14)00314-X/sref45http://refhub.elsevier.com/S1875-5100(14)00314-X/sref46http://refhub.elsevier.com/S1875-5100(14)00314-X/sref46http://refhub.elsevier.com/S1875-5100(14)00314-X/sref46http://refhub.elsevier.com/S1875-5100(14)00314-X/sref46http://refhub.elsevier.com/S1875-5100(14)00314-X/sref46http://refhub.elsevier.com/S1875-5100(14)00314-X/sref47http://refhub.elsevier.com/S1875-5100(14)00314-X/sref47http://refhub.elsevier.com/S1875-5100(14)00314-X/sref47http://refhub.elsevier.com/S1875-5100(14)00314-X/sref47http://refhub.elsevier.com/S1875-5100(14)00314-X/sref46http://refhub.elsevier.com/S1875-5100(14)00314-X/sref46http://refhub.elsevier.com/S1875-5100(14)00314-X/sref46http://refhub.elsevier.com/S1875-5100(14)00314-X/sref46http://refhub.elsevier.com/S1875-5100(14)00314-X/sref45http://refhub.elsevier.com/S1875-5100(14)00314-X/sref45http://refhub.elsevier.com/S1875-5100(14)00314-X/sref45http://refhub.elsevier.com/S1875-5100(14)00314-X/sref45http://refhub.elsevier.com/S1875-5100(14)00314-X/sref44http://refhub.elsevier.com/S1875-5100(14)00314-X/sref44http://refhub.elsevier.com/S1875-5100(14)00314-X/sref44http://refhub.elsevier.com/S1875-5100(14)00314-X/sref44http://refhub.elsevier.com/S1875-5100(14)00314-X/sref44http://refhub.elsevier.com/S1875-5100(14)00314-X/sref43http://refhub.elsevier.com/S1875-5100(14)00314-X/sref43http://refhub.elsevier.com/S1875-5100(14)00314-X/sref43http://refhub.elsevier.com/S1875-5100(14)00314-X/sref43http://refhub.elsevier.com/S1875-5100(14)00314-X/sref42http://refhub.elsevier.com/S1875-5100(14)00314-X/sref42http://refhub.elsevier.com/S1875-5100(14)00314-X/sref42http://refhub.elsevier.com/S1875-5100(14)00314-X/sref41http://refhub.elsevier.com/S1875-5100(14)00314-X/sref41http://refhub.elsevier.com/S1875-5100(14)00314-X/sref40http://refhub.elsevier.com/S1875-5100(14)00314-X/sref40http://refhub.elsevier.com/S1875-5100(14)00314-X/sref40http://refhub.elsevier.com/S1875-5100(14)00314-X/sref39http://refhub.elsevier.com/S1875-5100(14)00314-X/sref39http://refhub.elsevier.com/S1875-5100(14)00314-X/sref39http://refhub.elsevier.com/S1875-5100(14)00314-X/sref39http://refhub.elsevier.com/S1875-5100(14)00314-X/sref39http://www.gly.uga.edu/railsback/PGSG/PGSGmain.htmlhttp://www.gly.uga.edu/railsback/PGSG/PGSGmain.htmlhttp://refhub.elsevier.com/S1875-5100(14)00314-X/sref37http://refhub.elsevier.com/S1875-5100(14)00314-X/sref37http://refhub.elsevier.com/S1875-5100(14)00314-X/sref37http://refhub.elsevier.com/S1875-5100(14)00314-X/sref37http://refhub.elsevier.com/S1875-5100(14)00314-X/sref37http://refhub.elsevier.com/S1875-5100(14)00314-X/sref37http://refhub.elsevier.com/S1875-5100(14)00314-X/sref36http://refhub.elsevier.com/S1875-5100(14)00314-X/sref36http://refhub.elsevier.com/S1875-5100(14)00314-X/sref36http://refhub.elsevier.com/S1875-5100(14)00314-X/sref36http://refhub.elsevier.com/S1875-5100(14)00314-X/sref35http://refhub.elsevier.com/S1875-5100(14)00314-X/sref35http://refhub.elsevier.com/S1875-5100(14)00314-X/sref34http://refhub.elsevier.com/S1875-5100(14)00314-X/sref34http://refhub.elsevier.com/S1875-5100(14)00314-X/sref34http://refhub.elsevier.com/S1875-5100(14)00314-X/sref34http://refhub.elsevier.com/S1875-5100(14)00314-X/sref34http://refhub.elsevier.com/S1875-5100(14)00314-X/sref33http://refhub.elsevier.com/S1875-5100(14)00314-X/sref33http://refhub.elsevier.com/S1875-5100(14)00314-X/sref33http://refhub.elsevier.com/S1875-5100(14)00314-X/sref33http://refhub.elsevier.com/S1875-5100(14)00314-X/sref33http://refhub.elsevier.com/S1875-5100(14)00314-X/sref32http://refhub.elsevier.com/S1875-5100(14)00314-X/sref32http://refhub.elsevier.com/S1875-5100(14)00314-X/sref32http://refhub.elsevier.com/S1875-5100(14)00314-X/sref32http://refhub.elsevier.com/S1875-5100(14)00314-X/sref31http://refhub.elsevier.com/S1875-5100(14)00314-X/sref31http://refhub.elsevier.com/S1875-5100(14)00314-X/sref31http://refhub.elsevier.com/S1875-5100(14)00314-X/sref30http://refhub.elsevier.com/S1875-5100(14)00314-X/sref30http://refhub.elsevier.com/S1875-5100(14)00314-X/sref30http://refhub.elsevier.com/S1875-5100(14)00314-X/sref30http://refhub.elsevier.com/S1875-5100(14)00314-X/sref30http://refhub.elsevier.com/S1875-5100(14)00314-X/sref29http://refhub.elsevier.com/S1875-5100(14)00314-X/sref29http://refhub.elsevier.com/S1875-5100(14)00314-X/sref29http://refhub.elsevier.com/S1875-5100(14)00314-X/sref29http://refhub.elsevier.com/S1875-5100(14)00314-X/sref29http://refhub.elsevier.com/S1875-5100(14)00314-X/sref28http://refhub.elsevier.com/S1875-5100(14)00314-X/sref28http://refhub.elsevier.com/S1875-5100(14)00314-X/sref28http://refhub.elsevier.com/S1875-5100(14)00314-X/sref27http://refhub.elsevier.com/S1875-5100(14)00314-X/sref27http://refhub.elsevier.com/S1875-5100(14)00314-X/sref26http://refhub.elsevier.com/S1875-5100(14)00314-X/sref26http://refhub.elsevier.com/S1875-5100(14)00314-X/sref25http://refhub.elsevier.com/S1875-5100(14)00314-X/sref25http://refhub.elsevier.com/S1875-5100(14)00314-X/sref25http://refhub.elsevier.com/S1875-5100(14)00314-X/sref25http://refhub.elsevier.com/S1875-5100(14)00314-X/sref24http://refhub.elsevier.com/S1875-5100(14)00314-X/sref24http://refhub.elsevier.com/S1875-5100(14)00314-X/sref24http://refhub.elsevier.com/S1875-5100(14)00314-X/sref23http://refhub.elsevier.com/S1875-5100(14)00314-X/sref23http://refhub.elsevier.com/S1875-5100(14)00314-X/sref23http://refhub.elsevier.com/S1875-5100(14)00314-X/sref22http://refhub.elsevier.com/S1875-5100(14)00314-X/sref22http://refhub.elsevier.com/S1875-5100(14)00314-X/sref22http://refhub.elsevier.com/S1875-5100(14)00314-X/sref22http://refhub.elsevier.com/S1875-5100(14)00314-X/sref21http://refhub.elsevier.com/S1875-5100(14)00314-X/sref21http://refhub.elsevier.com/S1875-5100(14)00314-X/sref21http://refhub.elsevier.com/S1875-5100(14)00314-X/sref20http://refhub.elsevier.com/S1875-5100(14)00314-X/sref20http://refhub.elsevier.com/S1875-5100(14)00314-X/sref20http://refhub.elsevier.com/S1875-5100(14)00314-X/sref19http://refhub.elsevier.com/S1875-5100(14)00314-X/sref19http://refhub.elsevier.com/S1875-5100(14)00314-X/sref19http://refhub.elsevier.com/S1875-5100(14)00314-X/sref19http://refhub.elsevier.com/S1875-5100(14)00314-X/sref19http://refhub.elsevier.com/S1875-5100(14)00314-X/sref18http://refhub.elsevier.com/S1875-5100(14)00314-X/sref18http://refhub.elsevier.com/S1875-5100(14)00314-X/sref17http://refhub.elsevier.com/S1875-5100(14)00314-X/sref17http://refhub.elsevier.com/S1875-5100(14)00314-X/sref16http://refhub.elsevier.com/S1875-5100(14)00314-X/sref16http://refhub.elsevier.com/S1875-5100(14)00314-X/sref16http://refhub.elsevier.com/S1875-5100(14)00314-X/sref16http://refhub.elsevier.com/S1875-5100(14)00314-X/sref16http://refhub.elsevier.com/S1875-5100(14)00314-X/sref15http://refhub.elsevier.com/S1875-5100(14)00314-X/sref15http://refhub.elsevier.com/S1875-5100(14)00314-X/sref15http://refhub.elsevier.com/S1875-5100(14)00314-X/sref15http://refhub.elsevier.com/S1875-5100(14)00314-X/sref14http://refhub.elsevier.com/S1875-5100(14)00314-X/sref14http://refhub.elsevier.com/S1875-5100(14)00314-X/sref14http://refhub.elsevier.com/S1875-5100(14)00314-X/sref13http://refhub.elsevier.com/S1875-5100(14)00314-X/sref13http://refhub.elsevier.com/S1875-5100(14)00314-X/sref12http://refhub.elsevier.com/S1875-5100(14)00314-X/sref12http://refhub.elsevier.com/S1875-5100(14)00314-X/sref12http://refhub.elsevier.com/S1875-5100(14)00314-X/sref12http://refhub.elsevier.com/S1875-5100(14)00314-X/sref12http://refhub.elsevier.com/S1875-5100(14)00314-X/sref11http://refhub.elsevier.com/S1875-5100(14)00314-X/sref11http://refhub.elsevier.com/S1875-5100(14)00314-X/sref11http://refhub.elsevier.com/S1875-5100(14)00314-X/sref11http://refhub.elsevier.com/S1875-5100(14)00314-X/sref11http://refhub.elsevier.com/S1875-5100(14)00314-X/sref10http://refhub.elsevier.com/S1875-5100(14)00314-X/sref10http://refhub.elsevier.com/S1875-5100(14)00314-X/sref10http://refhub.elsevier.com/S1875-5100(14)00314-X/sref10http://refhub.elsevier.com/S1875-5100(14)00314-X/sref9http://refhub.elsevier.com/S1875-5100(14)00314-X/sref9http://refhub.elsevier.com/S1875-5100(14)00314-X/sref8http://refhub.elsevier.com/S1875-5100(14)00314-X/sref8http://refhub.elsevier.com/S1875-5100(14)00314-X/sref8http://refhub.elsevier.com/S1875-5100(14)00314-X/sref7http://refhub.elsevier.com/S1875-5100(14)00314-X/sref7http://refhub.elsevier.com/S1875-5100(14)00314-X/sref7http://refhub.elsevier.com/S1875-5100(14)00314-X/sref6http://refhub.elsevier.com/S1875-5100(14)00314-X/sref6http://refhub.elsevier.com/S1875-5100(14)00314-X/sref6http://refhub.elsevier.com/S1875-5100(14)00314-X/sref6http://refhub.elsevier.com/S1875-5100(14)00314-X/sref4http://refhub.elsevier.com/S1875-5100(14)00314-X/sref4http://refhub.elsevier.com/S1875-5100(14)00314-X/sref4http://refhub.elsevier.com/S1875-5100(14)00314-X/sref4http://refhub.elsevier.com/S1875-5100(14)00314-X/sref4http://refhub.elsevier.com/S1875-5100(14)00314-X/sref3http://refhub.elsevier.com/S1875-5100(14)00314-X/sref3http://refhub.elsevier.com/S1875-5100(14)00314-X/sref3http://refhub.elsevier.com/S1875-5100(14)00314-X/sref3http://refhub.elsevier.com/S1875-5100(14)00314-X/sref3http://refhub.elsevier.com/S1875-5100(14)00314-X/sref2http://refhub.elsevier.com/S1875-5100(14)00314-X/sref2http://refhub.elsevier.com/S1875-5100(14)00314-X/sref2http://refhub.elsevier.com/S1875-5100(14)00314-X/sref2http://refhub.elsevier.com/S1875-5100(14)00314-X/sref1http://refhub.elsevier.com/S1875-5100(14)00314-X/sref1

  • 8/9/2019 1-s2.0-S187551001400314X-main

    12/12

    Wang, H., Marongiu-Porcu, M., Economides, M.J., 2014. Poroelastic versus poro-plastic modeling of hydraulic fracturing. In: SPE Paper 168,600 Presented at theSPE Hydraulic Fracturing Technology Conference Held in Woodlands, Texas,USA, 4e6 February.

    Yahya, N., Kashif, M., Nasir, N., Akhtar, M.N., Yusof, N.M., 2012. Cobalt ferritenanoparticles: an innovative approach for enhanced oil recovery application.

     J. Nano Res. 17, 115e126.

    Zhang, T., Ellis, G.E., Ruppel, S.C., Milliken, K., Lewan, M., Sun, X., 2013. Effect of organic matter properties, clay mineral type and thermal maturity on gasadsorption in organic-rich shale systems. In: Paper was Presented at Uncon-ventional Resources Technology Conference, Held at Denver, Colorado, USA,August 12e14.

    H. Wang et al. / Journal of Natural Gas Science and Engineering 21 (2014) 874e885   885

    http://refhub.elsevier.com/S1875-5100(14)00314-X/sref48http://refhub.elsevier.com/S1875-5100(14)00314-X/sref48http://refhub.elsevier.com/S1875-5100(14)00314-X/sref48http://refhub.elsevier.com/S1875-5100(14)00314-X/sref48http://refhub.elsevier.com/S1875-5100(14)00314-X/sref48http://refhub.elsevier.com/S1875-5100(14)00314-X/sref48http://refhub.elsevier.com/S1875-5100(14)00314-X/sref49http://refhub.elsevier.com/S1875-5100(14)00314-X/sref49http://refhub.elsevier.com/S1875-5100(14)00314-X/sref49http://refhub.elsevier.com/S1875-5100(14)00314-X/sref49http://refhub.elsevier.com/S1875-5100(14)00314-X/sref50http://refhub.elsevier.com/S1875-5100(14)00314-X/sref50http://refhub.elsevier.com/S1875-5100(14)00314-X/sref50http://refhub.elsevier.com/S1875-5100(14)00314-X/sref50http://refhub.elsevier.com/S1875-5100(14)00314-X/sref50http://refhub.elsevier.com/S1875-5100(14)00314-X/sref50http://refhub.elsevier.com/S1875-5100(14)00314-X/sref50http://refhub.elsevier.com/S1875-5100(14)00314-X/sref50http://refhub.elsevier.com/S1875-5100(14)00314-X/sref50http://refhub.elsevier.com/S1875-5100(14)00314-X/sref50http://refhub.elsevier.com/S1875-5100(14)00314-X/sref50http://refhub.elsevier.com/S1875-5100(14)00314-X/sref50http://refhub.elsevier.com/S1875-5100(14)00314-X/sref49http://refhub.elsevier.com/S1875-5100(14)00314-X/sref49http://refhub.elsevier.com/S1875-5100(14)00314-X/sref49http://refhub.elsevier.com/S1875-5100(14)00314-X/sref49http://refhub.elsevier.com/S1875-5100(14)00314-X/sref48http://refhub.elsevier.com/S1875-5100(14)00314-X/sref48http://refhub.elsevier.com/S1875-5100(14)00314-X/sref48http://refhub.elsevier.com/S1875-5100(14)00314-X/sref48http://refhub.elsevier.com/S1875-5100(14)00314-X/sref48