Fracture & Matrix Kr Analysis

19
 Simulation of counter-current imbibition in water-wet fractured reservoirs Hassan Sh. Behbahani, Ginevra Di Donato, Martin J. Blunt  *  Departmen t of Earth Science and Engineerin g, Imperial College London, SW7 2AZ, UK Received 26 November 2003; received in revised form 5 August 2005; accepted 9 August 2005 Abstract Counter-current imbibition, where water spontaneously enters a water-wet rock while oil escapes by flowing in the opposite direction is a key recovery mechanism in fractured reservoirs. Fine grid, one- and two-dimensional simulations of counter-current imbibition are performed and the results are compared with experimental measurements in the literature. The experimental data are reproduced using two sets of relative permeabilities and capillary pressures. One set is derived from a counter-current experiment and one set is computed using pore-scale modelling. Two-dimensional simulations of water flow through a single high permeability fracture in contact with a lower permeability matrix are run. The results are compared with experimental measurements of fracture flow and matrix imbibition in the literature and with one-dimensional simulations that account for imbibition from fracture to matrix using an empirical transfer function. It is shown that with this transfer function the behavior of the two-dimensional displacement can be predicted using a one-dimensional model. D 2005 Published by Elsevier B.V.  Keywor ds:  Counter-current imbibition; Dual porosity models; Fracture/matrix transfer function; Network modelling 1. Introduction Fr act ur ed carbonate reservoirs are impor tant oil and gas resour ces. These reservo ir s ar e compose d of two con tinua: the fr acture net wor k and matr ix. The fractures typi call y have a high per mea bili ty but ve ry low volume compar ed to the ma tr ix, whos e  permeability may be several orders of magnitude lower, but which conta ins the ma jor ity of the reco- verable oil. Waterfl ood ing is fr equ ent ly imple men ted to in- crease recovery in fractured reservoirs. However, the  performance of waterflooding depends crucially on the wetta bi lit y of the reservo ir . If the re servo ir is oi l- wet, water will not re adily di splace oil in the mat rix and only the oil in the fra ct ur es will be di splace d, resulting in poor re coveri es and ea rl y water breakthrough. In water-wet fractured reservoirs, imbibition can lead to sign ific ant rec over ies. Imbibi- tion is the mechanism of displacement of non-wetting  phase by wetting phase. Strong capillary forces lead to the imbibition of water as the wett ing phase int o the matrix and the discharged oil is displaced into the fractures. Imbib ition can tak e pla ce by co- current and/or  count er -cur rent flow (Pa rs ons and Chaney , 1966; Iffly et al., 1972; Hamon and Vidal, 1986; Bourbiaux and Kalay dji an, 1990; Al -Lawati and Saleh , 199 6; 0920-4105/$ - see front matter  D 2005 Published by Elsevier B.V. doi:10.1016/j.petrol.2005.08.001 * Corresponding author. Tel.: +44 20 7594 6500; fax: +44 20 7594 7444.  E-mail address:  [email protected] (M.J. Blunt). Journal of Petroleum Science and Engineering 50 (2006) 21–39 www.elsevier.com/locate/petrol

description

RELATIVE PERMEABILITY

Transcript of Fracture & Matrix Kr Analysis

  • ren

    re

    ra

    ring,

    ised f

    1. Introduction

    crease recovery in fractured reservoirs. However, the

    performance of waterflooding depends crucially on

    fractures.

    Imbibition can take place by co-current and/or

    nd Chaney, 1966;

    Journal of Petroleum Science and EnginFractured carbonate reservoirs are important oil

    and gas resources. These reservoirs are composed

    of two continua: the fracture network and matrix.

    The fractures typically have a high permeability but

    very low volume compared to the matrix, whose

    permeability may be several orders of magnitude

    lower, but which contains the majority of the reco-

    verable oil.

    Waterflooding is frequently implemented to in-

    the wettability of the reservoir. If the reservoir is

    oil-wet, water will not readily displace oil in the

    matrix and only the oil in the fractures will be

    displaced, resulting in poor recoveries and early

    water breakthrough. In water-wet fractured reservoirs,

    imbibition can lead to significant recoveries. Imbibi-

    tion is the mechanism of displacement of non-wetting

    phase by wetting phase. Strong capillary forces lead

    to the imbibition of water as the wetting phase into

    the matrix and the discharged oil is displaced into theAbstract

    Counter-current imbibition, where water spontaneously enters a water-wet rock while oil escapes by flowing in the opposite

    direction is a key recovery mechanism in fractured reservoirs. Fine grid, one- and two-dimensional simulations of counter-current

    imbibition are performed and the results are compared with experimental measurements in the literature. The experimental data are

    reproduced using two sets of relative permeabilities and capillary pressures. One set is derived from a counter-current experiment

    and one set is computed using pore-scale modelling.

    Two-dimensional simulations of water flow through a single high permeability fracture in contact with a lower permeability

    matrix are run. The results are compared with experimental measurements of fracture flow and matrix imbibition in the literature

    and with one-dimensional simulations that account for imbibition from fracture to matrix using an empirical transfer function. It is

    shown that with this transfer function the behavior of the two-dimensional displacement can be predicted using a one-dimensional

    model.

    D 2005 Published by Elsevier B.V.

    Keywords: Counter-current imbibition; Dual porosity models; Fracture/matrix transfer function; Network modellingSimulation of counter-cur

    fractured

    Hassan Sh. Behbahani, Ginev

    Department of Earth Science and Enginee

    Received 26 November 2003; received in rev0920-4105/$ - see front matter D 2005 Published by Elsevier B.V.

    doi:10.1016/j.petrol.2005.08.001

    * Correspondi

    7444.

    E-mail address: [email protected] (M.J. Blunt).t imbibition in water-wet

    servoirs

    Di Donato, Martin J. Blunt *

    Imperial College London, SW7 2AZ, UK

    orm 5 August 2005; accepted 9 August 2005

    eering 50 (2006) 2139

    www.elsevier.com/locate/petrolal, 1986; Bourbiauxcounter-current flow (Parsons a

    Iffly et al., 1972; Hamon and Vidng author. Tel.: +44 20 7594 6500; fax: +44 20 7594and Kalaydjian, 1990; Al-Lawati and Saleh, 1996;

  • eum SPooladi-Darvish and Firoozabadi, 2000). In co-current

    flow the water and oil flow in the same direction, and

    water pushes oil out of the matrix. In counter-current

    flow, the oil and water flow in opposite directions,

    and oil escapes by flowing back along the same

    direction along which water has imbibed. Co-current

    imbibition is faster and can be more efficient than

    counter-current imbibition (Bourbiaux and Kalaydjian,

    1990; Chimienti et al., 1999; Pooladi-Darvish and

    Firoozabadi, 2000) but counter-current imbibition is

    often the only possible displacement mechanism for

    cases where a region of the matrix is completely

    surrounded by water in the fractures (Pooladi-Darvish

    and Firoozabadi, 2000; Najurieta et al., 2001; Tang

    and Firoozabadi, 2001). Experimentally this process

    can be studied by surrounding a core sample with

    water and measuring the oil recovery as a function of

    time (Iffly et al., 1972; Du Prey, 1978; Hamon and

    Vidal, 1986; Bourbiaux and Kalaydjian, 1990; Cuiec

    et al., 1994; Zhang et al., 1996; Cil et al., 1998;

    Chimienti et al., 1999; Rangel-German and Kovscek,

    2002). The imbibition rate is controlled by the per-

    meability of the matrix, its porosity, the oil/water

    interfacial tension and the flow geometry (Mattax

    and Kyte, 1962; Iffly et al., 1972; Hamon and

    Vidal, 1986; Babadagli and Ershaghi, 1992; Al-Lawati

    and Saleh, 1996; Ma et al., 1997; Cil et al., 1998;

    Chimienti et al., 1999) although the ultimate recovery

    is generally only governed by the residual oil satura-

    tion in strongly water-wet systems.

    Correlations have been developed to predict the

    recovery from counter-current imbibition as a func-

    tion of time for different samples. Mattax and Kyte

    (1962) hypothesized that the oil recovery for sys-

    tems of different size, shape and fluid properties

    was a unique function of a dimensionless time.

    Ma et al. (1997) modified an expression derived

    by Mattax and Kyte (1962) to include the effect

    of the non-wetting phase viscosity. Their experimen-

    tal results showed that the imbibition time is in-

    versely proportional to the geometric mean of the

    water and oil viscosities. They proposed the follow-

    ing correlation:

    tD tK

    /

    sr

    lwlop 1

    L2c1

    where t is time, K is permeability, / is porosity, ris interfacial tension, lw and lo are water and oilviscosities and L is the characteristic length that is

    H.Sh. Behbahani et al. / Journal of Petrol22c

    determined by the size, shape, and boundary condi-tions of the sample and is defined by Zhang et al.

    (1996) as:

    Fc 1Vma

    Xs

    Ama

    lma2

    Lc 1

    Fc

    r3

    where Vma is the bulk volume of the matrix (core

    sample), Ama is the area of a surface open to flow in

    the flow direction, lma is the distance from the open

    surface to the no flow boundary and the summation is

    over all open surfaces of the block.

    Ma et al. (1997) showed that recovery as a function

    of time for a variety of experiments on different water-

    wet samples fall on a single universal curve as a func-

    tion of the dimensionless time, tD, Eq. (1). In particular,

    imbibition experimental data presented by Mattax and

    Kyte (1962) for Alundum samples and Weiler sand-

    stones, Hamon and Vidals (1986) results for synthetic

    materials and Zhang et al.s (1996) results for Berea

    sandstones with different boundary conditions all

    scaled onto the same curve that was reasonably well

    fitted by the following empirical function first proposed

    by Aronofsky et al. (1958):

    R Rl 1 eatD 4

    where R is the recovery, Rl is the ultimate recovery

    and a is a constant that best matches the data with avalue of approximately 0.05.

    Eq. (1) was proposed for strongly water-wet media

    and ignores the effects of wettability. Gupta and

    Civan (1994) and Cil et al. (1998) extended the

    definition of dimensionless time, Eq. (1), by including

    a cosh term, where h is the oil/water contact angle torepresent core wettability. For water-wet systems it is

    assumed that cosh =1. Omitting the impact of wetta-bility in Eq. (1) means that the coefficient a in Eq. (4)can also be a function of contact angle distribution.

    Furthermore, other authors have presented extensions

    to Eq. (4) that account for transfer from matrix to

    fracture and transfer from dead-end pores that better

    match experimental data (see, for instance, Civan,

    1994; Gupta and Civan, 1994; Viksund et al., 1998;

    Terez and Firoozabadi, 1999; and Matejka et al.,

    2002).

    Zhou et al. (2002) correlated the results of counter-

    current imbibition experiments on water-wet samples of

    cience and Engineering 50 (2006) 2139a diatomite outcrop with very high porosity and low

  • q. (1)

    ; Zhan

    H.Sh. Behbahani et al. / Journal of Petroleum Science and Engineering 50 (2006) 2139 23permeability for a wide range of mobility ratios with the

    scaling relation:

    tD tK

    /

    srL2c

    k rw4 k ro4

    p 1M4

    p 1M4

    p5

    where kr*=kr* /l is defined as a characteristic mobilityand M*(=krw* /kro* ) is defined as a characteristic mobil-ity ratio. They used end-point relative permeabilities to

    Fig. 1. Normalized oil recovery as a function of dimensionless time, E

    Kyte, 1962; Hamon and Vidal, 1986; Bourbiaux and Kalaydjian, 1990

    empirical fit to the data by Ma et al. (1997), Eq. (4).calculate kr* and M*.Fig. 1 shows experimental imbibition data in the

    literature. While the data considered by Ma et al.

    (1997) does approximately lie on a single curve

    given by Eq. (4), consideration of other results on

    water-wet samples from Bourbiaux and Kalaydjian

    (1990), Cil et al. (1998) and Zhou et al. (2000) shows

    more scatter in particular for some of the experi-

    ments the imbibition rate is much slower than predicted

    by Eq. (4). The experiments analyzed by Ma et al.

    Fig. 2. Grid system for 1D simulations of counter-curren(1997) were performed on either artificial porous

    media (that is, not rock samples) or with an initial

    water saturation of zero. The other data presented

    (Bourbiaux and Kalaydjian, 1990; Cil et al., 1998;

    Zhou et al., 2000) are all performed on sandstone

    cores with an initial water saturation present, which is

    more likely to be representative of reservoir displace-

    ments. Only the data for a nominal aging time of zero

    from Zhou et al.s (2000) experiments were used to

    , for counter-current imbibition on different core samples (Mattax and

    g et al., 1996; Cil et al., 1998; Zhou et al., 2000). The solid line is ancompare with other water-wet imbibition experimental

    results in Fig. 1, although in this case crude oil is the

    oleic phase and the system may not be strongly water-

    wet. It is possible that the scatter in this data comes

    from ignoring wettability in Eq. (1). However, while

    representing wettability in terms of cosh is appealing,the assignment of a single effective average contact

    angle is not physically correct for systems where

    there is a distribution of contact angles (Jackson et

    al., 2003; Behbahani and Blunt, 2004).

    t imbibition. A total of 42 grid blocks were used.

  • Furthermore the behavior for different initial water

    saturations is a challenge. The presence of initial water

    saturation reduces capillary pressure but increases the

    mobility of invading water. The competition between

    capillary pressure and relative permeability determines

    the recovery rate. Viksund et al. (1998) demonstrated

    rom counter-current imbibition experiments by Bour-

    iaux and Kalaydjian (1990) were used (Table 2). They

    erformed co-current, counter-current and simultaneous

    o- and counter-current wateroil imbibition tests on

    trongly water-wet sandstone cores. They simulated

    ounter-current imbibition using a 1D numerical

    odel. They reduced the measured co-current relative

    ermeabilities by approximately 30% to match the

    able 2

    atrix rock and fluid properties used in 1D and 2D counter-current

    imulations

    arameter Unit Base models* Network

    modelling

    data**

    orosity frac 0.233 0.2

    ermeability m2 1241015 3.1481012il density Kg m3 760 835

    able 3

    racture and matrix dimensions in 1D and 2D counter-current simula-

    ons

    imension Unit 1D modelling 2D modelling

    Fracture Matrix Fracture Matrix

    cm 6.1 6.1 9 9

    cm 1.2 28 9 9

    cm 2.1 2.1 1 1

    H.Sh. Behbahani et al. / Journal of Petroleum Science and Engineering 50 (2006) 213924that the imbibition rate for low permeability Berea,

    shown in Fig. 1, is around eight times slower than for

    most other media. Baldwin and Spinler (2002) moni-

    tored saturation profiles during spontaneous counter-

    current imbibition using magnetic resonance imaging

    (MRI)they showed a transition from a flat frontal

    advance to a more gradual water encroachment as the

    initial water saturation was increased.

    2. Simulation of 1D-counter-current imbibition

    Counter-current imbibition in one dimension (1D)

    was modeled using EclipseR-100, which is an industry-standard reservoir simulator. In the 1D model, illustrat-

    ed in Fig. 2, a region initially full of water with fracture

    properties (Tables 1 and 3) was connected to a thin

    matrix slab (Tables 2 and 3). A total of 42 grid blocks

    was used in the simulation, with smaller blocks at the

    inlet to capture accurately the initial advance of water

    into the matrix. Sensitivity studies using 200 grid

    blocks demonstrated that sufficient grid blocks were

    used to obtain converged results. All other faces not

    in contact with the fracture were closed (no flow bound-

    aries). The relative permeabilities of the fractures were

    assumed to be linear functions of saturation with no

    irreducible or residual saturation and the capillary pres-

    sure in the fracture was zero. The matrix relative per-

    meabilities and capillary pressures are discussed later.

    Conservation of water volume assuming incom-

    pressible flow and Darcys law in one dimension with

    no total velocity can be expressed as follows:

    /BSw

    Bt BBx

    kwkokt

    KBPc

    BSw

    BSw

    Bx gx qw qo

    0

    6where Pc is the capillary pressure, the mobility k =kr /lwhere kr is the relative permeability and kt=kw+ko. q

    Table 1

    Fracture properties used in 1D and 2D counter-current simulations

    Parameter Unit Value

    Initial water saturation Fraction 1.0

    Porosity Fraction 1.0

    Capillary pressure Pa 0.0Permeability m2 501012f

    b

    p

    c

    s

    c

    m

    p

    T

    F

    ti

    D

    X

    Y

    Zis the density and gx is the component of gravity in the

    flow direction. The reservoir simulator solves Eq. (6)

    on the grid shown in Fig. 2 using an implicit finite

    difference formulation with upstream weighting of the

    mobility terms (Aziz and Settari, 1979). It is assumed

    that this conventional treatment of the flow equations

    with saturation-dependent capillary pressure, Pc, and

    relative permeabilities, kr, is sufficient to predict the

    results of the experiments. Recently it has been sug-

    gested by Barenblatt et al. (2003) that a fundamentally

    different formulation of the conservation equations with

    rate-dependent coefficients is necessary to explain the

    results of imbibition experiments.

    2.1. Base models

    For the base case, the matrix and fluid properties

    Oil viscosity Pa s 1.5103 39.25103***Water density Kg m3 1090 1010Water viscosity Pa s 1.2103 0.967103***Interfacial tension mN m1 35 30Initial water saturation Fraction 0.40 0.25

    Residual oil saturation Fraction 0.422 0.75

    *: From Bourbiaux and Kalaydjian (1990).

    **: From Jackson et al. (2003).

    ***: From Zhou et al. (2000).T

    M

    s

    P

    P

    P

    OLc cm 28 3.18 28 3.18

  • counter-current experiment results and concluded that biaux and Kalaydjian (1990) using the same relative

    Fig. 3. Matrix imbibition relative permeabilities and capillary pressure used in the base models from Bourbiaux and Kalaydjian (1990).

    H.Sh. Behbahani et al. / Journal of Petroleum Science and Engineering 50 (2006) 2139 25co-current relative permeability measurements cannot

    be used to simulate counter-current imbibition. Count-

    er-current relative permeabilities and imbibition capil-

    lary pressure from Bourbiaux and Kalaydjian (1990),

    Fig. 3, were used in all base models.

    Fig. 4 shows the measured and predicted oil recov-

    ery from Bourbiaux and Kalaydjians (1990) experi-

    ments. In the experiments, the core was held

    vertically, and so gravity effects are included in the

    simulation. The simulation results are identical to

    experiments and to simulations performed by Bour-Fig. 4. Comparison of experimental and simulated (crosses) recoveries for co

    Kalaydjian (1990). Also shown are the simulated results from Bourbiaux anpermeabilities. The relative permeabilities and capillary

    pressure used for the simulations were chosen by Bour-

    biaux and Kalaydjian (1990) to match their experimen-

    tal results and so this is not a genuine prediction.

    Fig. 5 shows comparisons of the simulation results

    with the experimental data in Fig. 1. In most of these

    experiments the cores were held horizontally, and so for

    comparison we show a simulation result where gravi-

    tational effects have been neglected. This makes little

    difference to the results, except at late time and slightly

    improves the match to some of the other data.unter-current imbibition. The experimental data is from Bourbiaux and

    d Kalaydjian (1990) that agree very well with our results.

  • sing

    the M

    H.Sh. Behbahani et al. / Journal of Petroleum Science and Engineering 50 (2006) 2139262.2. Experimental data predicted by network modelling

    The pore-scale model used a network derived from

    a sample of Berea sandstone and can accurately predict

    steady-state waterflood relative permeabilities for

    Berea (Blunt et al., 2002; Jackson et al., 2003; Val-

    vatne and Blunt, 2004). The void space of the rock was

    described by a network of pores connected by throats.

    The pores and throats were assigned some idealized

    Fig. 5. Comparison of 1D vertical and horizontal simulation results u

    Kalaydjian (1990) compared to experimental data in the literature andgeometry and rules were developed to determine the

    multiphase fluid configurations and transport in these

    elements. The rules were combined in the network to

    Fig. 6. Matrix waterflood relative permeabilities and capillary pressure derivecompute effective transport properties on a mesoscopic

    scale some tens of pores across. The appropriate pore-

    scale physics combined with a geologically represen-

    tative description of the pore space gives a model that

    can predict average behavior, such as capillary pressure

    and relative permeability (Blunt et al., 2002; Jackson et

    al., 2003; Valvatne and Blunt, 2004). The imbibition

    relative permeabilities and capillary pressure predicted

    for Berea by network modelling and used in the simu-

    relative permeability and capillary pressure data from Bourbiaux and

    a et al. (1997) expression, Eq. (4).lations are shown in Fig. 6.

    The experiments by Zhou et al. (2000) were per-

    formed on Berea sandstone. Hence, of all the results

    d from network modelling of Berea sandstone by Jackson et al. (2003).

  • el de

    ne.

    H.Sh. Behbahani et al. / Journal of Petroleum Science and Engineering 50 (2006) 2139 27shown in Fig. 1, it is expected that the network

    model properties best represent this experiment. For

    a more careful comparison the oil and water visco-

    sities from Zhou et al. (2000) were used in the

    simulations.

    Fig. 7 shows the results of 1D counter-current

    imbibition simulations using network model derived

    data. The simulations accurately predict Zhou et al.s

    (2000) experimental results. This is a genuine predic-

    Fig. 7. 1D counter-current imbibition simulations using network mod

    literature and the experiments of Zhou et al. (2000) on Berea sandstotion of the experimental results, in that the relative

    permeabilities and capillary pressure used were inde-

    Fig. 8. Grid system used for the 2D counter-current imbibitionpendently computed and no parameters were tuned to

    match the results. The predicted imbibition rate is

    approximately ten times slower than measured on

    artificial porous media or systems with no initial

    water saturation present, as shown experimentally by

    Viksund et al. (1998).

    3. Simulation of 2D-counter-current imbibition

    rived data shown in Fig. 6 compared to experimental results in theA series of 2D simulations were run to check the

    1D results. The grid system used is shown in Fig. 8.

    simulations. A total of 115,236 grid blocks were used.

  • Fig. 9. Comparison of 2D counter-current imbibition simulations using Bourbiaux and Kalaydjian (1990) (base case, Fig. 3) and network model

    properties (Fig. 6) with literature experimental data.

    H.Sh. Behbahani et al. / Journal of Petroleum Science and Engineering 50 (2006) 213928A total of 115,235 fine grid blocks were used.

    Sensitivity studies using 90,000 and 160,000 grid

    blocks demonstrated that sufficient grid blocks

    were used to obtain converged results. Fig. 9

    shows the results for both sets of relative perme-

    abilities. For the 2D simulations, including gravity

    had no effect on the results. Again the agreement

    with experiments is good, particularly at early times.

    The results are similar to those obtained in 1D, Figs.5 and 7.

    Fig. 10. Effect of different oil to water viscosity ratios,M, on the results of 1D

    as a function of the dimensionless time, tD, proposed by Ma et al. (1997), E

    validating the correlation.4. Validity of the correlations, Eqs. (1) and (5)

    The dimensionless time, Eq. (1), can be derived

    using dimensional analysis. However, the scaling with

    viscosity is not obvious. Mattax and Kyte (1962) sug-

    gested that tD should be inversely proportional to the

    water viscosity, while a recent analytical approach by

    Tavassoli et al. (2005) suggests that tD is inversely

    proportional to the oil viscosity. The scaling in Eq.(1) involving the geometric average of the oil and

    simulations using network model derived data. The results are plotted

    q. (1). Notice that all the plots approximately fall onto a single curve,

  • Fig. 11. Using only the oil viscosity in the definition of dimensionless time leads to a large scatter in the results from 1D simulations with different

    oil to water viscosity ratios, M, using network model derived data.

    H.Sh. Behbahani et al. / Journal of Petroleum Science and Engineering 50 (2006) 2139 29water viscosities was based on experimental results for

    systems with different viscosity ratio by Ma et al.

    (1997). Zhou et al.s (2002) scaling equation, Eq. (5),

    also accounts for the mobility of wetting and non-wet-

    ting phases. To check the validity of Eqs. (1) and (5), a

    series of 1D-simulations for different oil to water vis-

    cosity ratios were performed using the network model

    derived data. The experiments by Zhang et al. (1996)

    covered a range of viscosity ratios from 4 to 160 inFig. 12. Using only water viscosity in the definition of dimensionless time le

    oil to water viscosity ratios, M, using network model derived data.our simulations a range from 0.01 to 200 was used. The

    recovery from the simulations plotted as a function of

    the dimensionless time, Eq. (1), is shown in Fig. 10.

    While it is evident that all the recoveries do not fall

    exactly on the same universal curve, there is little scatter

    in the results. In contrast using either the oil viscosity

    (Fig. 11) or the water viscosity (Fig. 12) only in the

    definition of dimensionless time or Eq. (5) (Fig. 13)

    leads to a much larger scatter in the curves, verifyingads to a large scatter in the results from 1D simulations with different

  • scatte

    eum Sthat Eq. (1) is at least approximately correct as the

    scaling function. These results confirm previous analy-

    tical and experimental studies by Ruth et al. (2000) and

    Fig. 13. Using Zhou et al. (2002) dimensionless time, Eq. (5) leads to a

    ratios, M, using network model derived data.

    H.Sh. Behbahani et al. / Journal of Petrol30Li et al. (2003). Ruth et al. (2003) performed a nume-

    rical study of the effect of matrix shape on imbibition

    rate and confirmed the accuracy of the expression for

    Lc, Eqs. (1)(3).

    5. Theory and simulation of fracture flow and

    imbibition

    5.1. Previous experimental studies

    Rangel-German and Kovscek (2002) investigated

    the rate of fracture to matrix transfer and the pattern

    of wetting fluid imbibition using CT scanning. They

    studied an air/water system where water flowed in a

    constant aperture fracture along the side of the system,

    and imbibed into the lower permeability matrix.

    Experiments with different flow rates and fracture

    apertures illustrated two different flow regimes.

    When the flow in the fractures was slow, there was

    sufficient time for the matrix to become saturated with

    water and there was a frontal advance in both matrix

    and fracturethis was called the filling-fracture re-

    gime and the amount of water in the matrix increased

    linearly with time. At higher fracture flow rates, water

    in the fracture moved ahead of water in the matrix.This was called the instantly filled regime and the

    amount of water in the matrix increased with the

    square root of time.

    r in the results from 1D simulations with different oil to water viscosity

    cience and Engineering 50 (2006) 21395.2. Theory for dual porosity systems

    At the field scale, flow in fractured reservoirs is

    simulated using a dual porosity approach (Barenblatt

    and Zheltov, 1960; Warren and Root, 1963) where

    conceptually the reservoir is composed of two domains:

    a flowing fraction, representing the fractures; and a

    relatively stagnant matrix. Transfer of fluid between

    fracture and matrix is represented by an transfer func-

    tion. For incompressible flow of oil and water in 1D the

    conservation equations for water are (Kazemi et al.,

    1992; Di Donato et al., 2003):

    /fBSwf

    Bt mt Bfwf

    Bx gx BG

    Bx T 7

    /m BSwm

    Bt T 8

    where the subscript f stands for fracture and m for

    matrix. vt is the total velocity. fwf is the water fractional

    flow in the fractures ignoring gravity:

    fwf kwfkwf kof 9

  • lution integral to compute the matrix saturation. This is

    consistent with using:

    T b/m Swf 1 Somr Swmi Swm Swmi 15

    A similar expression was derived by Kazemi et al

    (1992) from de Swaans (1978) convolution integral

    eum Science and Engineering 50 (2006) 2139 31and:

    G qw qo Kfwfkof 10

    T is the transfer function that represents the rate at

    which water transfers from fracture to matrix.

    de Swaan (1978), Kazemi et al. (1992), Di Donato

    et al. (2003), Di Donato and Blunt (2004) and Huang

    et al. (2004) developed transfer functions that repro-

    duce the exponential functional formEq. (4)that

    matches imbibition experiments. If Swm is the average

    saturation in the matrix:

    R

    Rl Swm Swmi

    1 Somr Swmi 11

    where Swmi is the initial water saturation in the matrix

    and Somr is the corresponding residual oil saturation,

    then from Eq. (4):

    Swm Swmi 1 Somr Swmi 1 eatD 12Then from Eq. (12):

    /mBSwm

    Bt T a tD

    t/m 1 Somr Swm

    b/m 1 Somr Swm 13

    where the rate constant b is defined by Eq. (13).In the experiments reviewed previously the effective

    fracture saturation was held at 1. Di Donato et al.

    (2003) assumed that the transfer function is indepen-

    dent of the fracture saturation, as long as SwfN0. This isconsistent with assuming that the capillary pressure in

    the low permeability matrix is much higher than in the

    Table 4

    Fracture properties used in 2D fracture/matrix simulations

    Parameter Unit Value

    Initial water saturation fraction 0.0

    Porosity fraction 1.0

    Capillary pressure Pa 0.0

    Permeability m2 151012Water relative permeability krw=Sw

    2

    Oil relative permeability kro= (1Sw)2

    H.Sh. Behbahani et al. / Journal of Petrolfractures the driving force for imbibition is the matrix

    capillary pressure and imbibition continues until the

    matrix and fracture capillary pressures are equal

    when Swf=0 and Swm=1Sorm. Thus:T b/m 1 Somr Swm SwfN0 0 Swf 0 14

    The transfer function is a linear function of the matrix

    saturation.

    de Swaan (1978) derived a similar transfer function

    to Eq. (14) when Swf=1. For Swfb1 he used a convo-This formulation is correct if the large-scale fracture

    saturation is viewed as being an average of fully satu-

    rated fractures undergoing imbibition and completely

    dry fractures. However, for uniformly partially saturat-

    ed fractures, it implies that imbibition ceases when the

    fracture and matrix saturations are proportional to each

    other, implying similar values of the fracture and matrix

    capillary pressures. This is rarely correct, and certainly

    inconsistent with the basic premise of dual porosity

    models that there is a huge disparity in permeability

    between flowing and stagnant regions. This transfer

    function also is linearly dependent on saturation.

    Other authors have derived transfer functions for

    imbibition based on a more accurate match to experi-

    mental data than Eq. (4) (Civan, 1994; Civan and

    Rasmussen, 2001, 2003; Penuela et al., 2002; Terez

    and Firoozabadi, 1999). In this work we will use Eq.

    (14) for our transfer function since it is based on a

    simple one-parameter fit to experiments.

    5.3. Grid-based simulation of fracture/matrix flow

    The purpose of this section is to determine: (1) if the

    behavior observed experimentally by Rangel-German

    and Kovscek (2002) can be reproduced by simulation;

    and (2) if the same results can be obtained using a 1D

    simulation to solve Eqs. (6) and (7) with an appropriate

    transfer function. The simulation model consisted of a

    horizontal fracture (Tables 4 and 5)) initially filled with

    oil (Swf=0) connected from both sides to oil saturated

    matrix blocks (Swm=Swc). The fracture was defined

    explicitly as a high permeability region with a porosity

    of 1, and no transfer function was used (Fig. 14). There

    were 122 matrix grid blocks along the fracture. The

    matrix grid system in the direction perpendicular to the

    fracture was exactly the same as those used to simulate

    Table 5

    Dimensions of the model in 2D fracture/matrix simulations

    Dimension Unit Fracture Matrix

    (on either side

    of the fracture)

    X cm .7 28 (20 .3, 221)Y cm 244 244 (1222)

    Z cm 1 1.

    .

  • counter-current imbibition in 1D (Fig. 2) that matched

    the experimental results. The system was 244 cm long

    with a width of 56.7 cm. The fracture aperture was 0.7

    Quadratic relative permeabilities were defined in the

    fracture (Table 4)this meant that a shock developed

    in the 1D displacement in the fracture and reduced

    Fig. 14. Grid system used for the 2D fracture/matrix simulation. The grid perpendicular to the fracture direction is exactly as in Fig. 2the darker

    area indicates finer grids close to the fracture.

    H.Sh. Behbahani et al. / Journal of Petroleum Science and Engineering 50 (2006) 213932cm (Table 5). Water was injected at a constant rate via

    an injection well into the first grid block of the fracture

    and liquid (oil and water) was produced at a constant

    pressure from a producing well that was completed in

    the last fracture grid block.

    The matrix porosity, relative permeabilities and ma-

    trix connate and residual saturations come from Bour-

    biaux and Kalaydjian (1990) (Table 2). In order to

    produce a situation similar to low matrix permeability

    fractured reservoirs, the matrix absolute permeability

    was reduced from 124 to 1 md to ensure a huge

    disparity in permeability between fracture and matrix.

    The matrix capillary pressure had the same functional

    form as used by Bourbiaux and Kalaydjian (1990), but

    increased by the square root of the ratio of the exper-

    imental permeability to the model matrix permeability.

    The viscosities of the oil and water were 1.5103 Pas. Capillary pressure in the fracture was zero. The

    fracture permeability was 151012 m2 (15 Darcy).Fig. 15. Comparison of fracture water saturations for 1D simulations with

    solution (dotted line) is compared to numerical simulation results using a com

    (dashed line). The good agreement shows that simulations can accurately renumerical dispersion in the simulation results. While

    simple analytical expressions for the fracture properties

    were used, the matrix relative permeabilities and capil-

    lary pressure were based on experimental data. Two

    water injection rates of 20 and 2 cc/h were used.

    Two additional simulations were run for comparison

    purposes. The first had no matrix and an injection rate

    of 20 cc/h. The results were compared with an analyt-

    ical solution based on the BuckleyLeverett approach

    and 1D numerical solutions using the streamline code

    described by Di Donato et al. (2003). The second was

    where the matrix had a capillary pressure of zero. This

    test shows the effect of viscous forces on the recovery

    process.

    5.4. Analytical and 1D numerical solutions

    Fig. 15 shows a comparison of numerical solutions

    for a simulation with no matrix compared to the ana-no matrix present after 3 and 5 h. The BuckleyLeverett analytical

    mercial reservoir simulator (solid line) and a streamline-based model

    produce a 1D displacement.

  • lytic BuckleyLeverett solution (Dake, 1978). The

    agreement between the numerical and analytical results

    confirms that both Eclipse and the streamline code can

    accurately reproduce a 1D displacement.

    Fig. 16 shows the fracture saturation when there is

    no capillary pressure in the matrix. The results are

    compared to the analytical solution where there is no

    matrix. The agreement between the two simulations

    indicates that in this model viscous forces alone have

    little effect on the fracture/matrix transfer.

    The results of 2D simulations with a finite matrix

    capillary pressure are now compared with results

    from 1D dual porosity models. The linear transfer

    functions developed by Di Donato et al. (2003) and

    Kazemi et al. (1992) were used in a dual porosity

    model. In the dual porosity model the following rosity simulation is calculated using the following

    olid lines show the fracture saturation for flow with matrix present and no

    Table 6

    Data used in the 1D dual porosity simulations

    Parameter Unit Value

    Fracture porosity, /f fraction 0.012Matrix porosity, /m fraction 0.233Fracture permeability, Kf m

    2 151012Matrix permeability, Km m

    2 1015

    Matrix initial water saturation, Swmi fraction 0.40

    Matrix residual oil saturation, Somr fraction 0.422

    Water density, qw Kg m3 1090

    Oil density, qo Kg m3 760

    Water viscosity, lw Pa s 1.5103Oil viscosity, lo Pa s 1.5103Fracture/Matrix rate constant, b days1 0.08423Fracture water relative permeability Krwf=Sw

    2

    Fracture oil relative permeability krof= (1Sw)2

    H.Sh. Behbahani et al. / Journal of Petroleum Science and Engineering 50 (2006) 2139 33equations were used to define fracture and matrix

    porosities and the effective permeability of the sys-

    tem from the geometry of the 2D explicit fracture

    model:

    /f Wf

    Wf Wm 16

    /m /mWm

    Wf Wm 17

    Keff WfKf WmKmWf Wm 18

    where W and K are the width and permeability,

    respectively. The total velocity in the 1D dual po-

    Fig. 16. Numerical simulations of fracture flow after 2 and 4 h. The scapillary pressure. The dotted lines are BuckleyLeverett analytical solutions assuming a purely 1D displacement with no matrix. The agreemen

    between the two simulations indicate that viscous forces have little effect on the fracture/matrix transfer.equation:

    mt QWfH

    19

    where Q is the injection rate and Wf is the fracture

    aperture. Tables 6 and 7 give the data used in the

    dual porosity simulations.

    5.5. Analysis of the results

    Figs. 17 and 18 show the comparison of the simu-

    lated water saturation profiles in the fracture and matrix

    for different time-steps for an injection rate of 20 cc/h.

    Due to the high injection rate it is expected that the

    front moves very fast in the fracture and most imbibi-

    tion take places when the average water saturation in

    the fracture is nearly one. Since the amount of transfer

    is small at early time, the fracture saturation profiles fort

  • all three simulations are similar. The simulated matrix

    saturation profiles are different, however, with the dual

    porosity models predicting more transfer into the matrix

    than the 2D simulation. This is simply because the

    transfer functions reproduce the empirical fit, Eq. (4)

    that gives a higher recovery than the equivalent 1D

    times for an injection rate of 2 cc/h. In this case water

    advances in the fracture and matrix at comparable rates.

    In contrast to the previous case, the simulated fracture

    saturation profiles are not the same for the different

    models. The simulated matrix saturation profiles are

    also different. Fig. 5 reveals that at early time (tD of

    less than approximately 5, or a real time of approxi-

    mately 70 h) the simulations predict a higher recovery

    than Eq. (4), while for intermediate and late times

    (greater than around 70 h) the simulations predict

    lower recovery. In our 2D simulations this relates to

    the recovery (or water saturation) in the matrix. Hence

    we expect that at early times the dual porosity models

    will give lower recovery in the matrix and higher

    recovery at late times. This is evident in Fig. 21. If

    Table 7

    Equivalent injection data used in 1D dual porosity simulations

    Parameter Unit Value

    High injection

    rate case

    Low injection

    rate case

    2D injection rate, Q2D cc h1 20 2

    2D total velocity, vt m s1 7.94105 7.94106

    1D injection rate, Q1D m3 day1 4.6676104 4.6676105

    H.Sh. Behbahani et al. / Journal of Petroleum Science and Engineering 50 (2006) 213934simulations. The results for 240 h correspond to a

    dimensionless time tD of 17. It is clear from a com-

    parison of Figs. 1 and 5 that at this time the simula-

    tions do indeed predict a lower recovery than the

    experimental correlation, Eq. (4), for 1D counter-cur-

    rent imbibition. The transfer function of Di Donato et

    al. (2003) shows little change in matrix saturation with

    distance, similar to the 2D simulations, while the

    Kazemi et al. (1992) function predicts a significant

    change in saturation with distance. This is because the

    Kazemi et al. (1992) model predicts less transfer into

    the matrix when the fracture saturation is less than

    1that is near the outlet.

    Fig. 19 shows the average matrix saturation as a

    function of time. The average saturation scales with

    the square root of time, except at late time, which was

    also observed experimentally for the instantly filled

    fracture regime by Rangel-German and Kovscek (2002).

    Figs. 20 and 21 compare the simulated water satu-

    ration profiles in the fracture and matrix at differentFig. 17. Simulated fracture water saturation for Q =20 cc/h after 3, 5 and 7 h

    using a transfer function due to either Di Donato et al. (2003) (dashed linethere is less imbibition into the matrix, more water is

    retained in the fracture and the waterfront in the fracture

    moves faster. This is confirmed in Fig. 20 where the

    water saturation in the fracture is moving faster for the

    dual porosity models as a consequence of the lower

    matrix recoveries. The water advance is still greater at

    70 h, which is just in the late time regime. However,

    most of the matrix has been next to a water-filled

    portion of the fracture for considerably less than 70 h,

    so the behavior is still consistent with the early time

    behavior.

    Another possible explanation for the discrepancy in

    the results is that the 2D simulations allow co-current

    flow, whereas the transfer functions are based on count-

    er-current flow, which is slower. In the 2D simulations

    oil can travel in a direction parallel to water towards the

    high permeability fracture and/or un-drained matrix

    grid blocks, resulting in more rapid recovery than

    from counter-current imbibition alone. However, a

    careful quantitative analysis of the results shows that. 2D simulation (solid line) is compared with a 1D dual porosity model

    ) or Kazemi et al. (1992) (dotted line).

  • Fig. 18. Simulated matrix water saturation for a flow rate of 20 cc/h after 100 and 240 h. 2D simulation (solid line) is compared with a dual porosity

    model using a transfer function due to either Di Donato et al. (2003) (dashed line) or Kazemi et al. (1992) (dotted line). The 2D simulation and the

    Di Donato et al. (2003) transfer function show a flat profile similar to that observed experimentally by Rangel-German and Kovscek (2002).

    H.Sh. Behbahani et al. / Journal of Petroleum Science and Engineering 50 (2006) 2139 35this does not affect recovery in this case, and that the

    differences between the results are explained solely by

    the differences in predicted 1D counter-current recov-

    ery. For instance, Fig. 21 shows that the water satura-

    tion at the inlet after 48 h is 43.5% in the 2D

    simulations and 42.8% using the Di Donato et al.

    (2003) model. 48 h corresponds to tD of 3.4. Using

    Fig. 5 to find the oil recovery assuming counter-current

    flow only for the simulations and using Eq. (4) will lead

    to exactly the same water saturations. This shows that

    co-current imbibition has no effect on the results. This

    is likely to be due to the extreme contrast in matrix and

    fracture permeabilities.

    The Kazemi et al. (1992) transfer function always

    gives a transfer rate that is less than or equal to thatFig. 19. The simulated average matrix water saturation in the 2D simulation s

    reproduces the behavior of the instantly filled regime in Rangel-German anpredicted by the Di Donato et al. (2003) model. As a

    consequence the Kazemi et al. (1992) model predicts

    less transfer into the matrix and a more diffuse and

    further advanced saturation profile in the fracture. Qual-

    itatively the 2D simulations give profiles more similar

    to the Di Donato et al. (2003) model, since in both cases

    a zero fracture capillary pressure is assumed.

    Overall the Di Donato et al. (2003) dual porosity

    model gives similar results to direct 2D simulation,

    indicating that a dual porosity approach can properly

    simulate fracture/matrix flow in a simple system. The

    differences in the results can all be explained by con-

    sidering the differences in the predictions of counter-

    current imbibition with no fracture flow. It should be

    possible to adjust the rate constant b to obtain a bettercales linearly with the square root of time for a flow rate of 20 cc/h and

    d Kovsceks (2002) experimental results.

  • Fig. 20. The fracture water saturation for a flow rate of 2 cc/h after 30 and 70 h. 2D simulation (solid line) is compared to dual porosity models using

    otted

    H.Sh. Behbahani et al. / Journal of Petroleum Science and Engineering 50 (2006) 213936match between the dual porosity and 2D simulation

    models, or to use a transfer function with more para-

    meters that better reproduce experimental data (Civan

    and Rasmussen, 2001).

    Fig. 22 shows the average matrix water saturation as

    a function of time. The average saturation scales ap-

    proximately linearly with time at early times (less than

    around 200 h), which also was observed experimentally

    for the filling-fracture regime by Rangel-German and

    Kovscek (2002).

    Penuela et al. (2002) developed a transfer function

    for dual porosity simulation that matched fine grid

    simulations of imbibition. They used a transfer function

    with the same definition of shape factor used conven-

    tionally by Kazemi et al. (1992), but allowed the shape

    factor to be time-dependent. In comparison our formu-

    lation is mathematically simpler, and physically more

    the Di Donato et al. (2003) (dashed line) and Kazemi et al. (1992) (dappealing, since the governing partial differential equa-

    Fig. 21. The matrix water saturation for a flow rate of 2 cc/h after 48 and

    porosity models using the Di Donato et al. (2003) (dashed line) and Kazemtions are written as a function of saturation. However

    incompressible two-phase flow is assumed in this

    work, whereas Penuela et al. (2002) also considered

    compressibility.

    Rangel-German and Kovscek (2003) also proposed

    a time-dependent shape factor to match their experi-

    mental results. They found a formulation that matched

    the overall recovery for different flow rates; however, in

    essence they upscaled a 2D system to a zero-dimen-

    sional average property. The dual porosity model pro-

    posed here upscales a 2D or 3D system into a 1D

    problem along the flow direction. Their formulation

    except for the representation of a time rather than a

    saturation dependent shape factoris broadly equiva-

    lent to Kazemi et al.s (1992) transfer function. When

    placed in a dual porosity model, this transfer function

    underestimates the matrix recovery and overestimates

    line) transfer functions.the water advance rate in the fractures. The formulation

    240 h. Results from 2D simulation (solid line) are compared to dual

    i et al. (1992) (dotted line) transfer functions.

  • ation

    ture r

    eum S(2) The validity of the Ma et al. (1997) scaling

    function, Eq. (1), was tested by performing simu-

    lations for different oil/water viscosity ratios

    using network model derived data (Jackson et

    al., 2003). The results indicated that recoveryin this paper is identical to theirs for a constant shape

    factor.

    6. Conclusions

    (1) Simulation of one-dimensional and two-dimen-

    sional counter-current imbibition matched the

    results of core experiments. This confirms that a

    conventional Darcy treatment of multiphase flow

    is adequate to describe counter-current imbibition.

    Fig. 22. The simulated average matrix water saturation in the 2D simul

    for a flow rate of 2 cc/h and reproduces the behavior of the filling-frac

    H.Sh. Behbahani et al. / Journal of Petrolplots for different viscosity ratios could be scaled

    onto the same curve using a dimensionless time

    that is inversely proportional to the geometric

    mean of the water and oil viscosities, as estab-

    lished experimentally by Ma et al. (1997).

    (3) The results of 2D simulations of flow in a long

    fracture connected to a horizontal water-wet ma-

    trix showed the same flow regimes as observed

    experimentally by Rangel-German and Kovscek

    (2002).

    (4) A 1D dual porosity model with a transfer function

    that matched core-scale imbibition experiments

    was able to reproduce the behavior observed

    using explicit 2D simulation of fracture/matrix

    flow.

    Nomenclature

    Ama area of a surface open to flow in the flow

    direction, m2Fc shape factor, m2

    fj fractional flow

    H vertical thickness, m

    K absolute permeability, m2 or Am2

    kr relative permeability

    L, Lc characteristic length, m

    lma distance from the open surface to the no flow

    boundary, m

    M viscosity ratio

    M* mobility ratio

    P pressure, Pa

    Q flow rate, m3 s1

    Rl final oil recovery, m3

    R recovery, m3

    S saturation, fraction

    scales approximately linearly with time at early times (less than 200 h)

    egime in Rangel-German and Kovsceks (2002) experimental results.

    cience and Engineering 50 (2006) 2139 37T transfer function, s1

    t time, s

    tD dimensionless time

    Vma bulk volume of matrix (core sample), m3

    Vt total fluid velocity, m s1

    W width, m

    Greek

    b rate constant, s1

    / porosity, fractionk mobility, Pa1 s1

    lo oil viscosity, cp, Pa slw water viscosity, cp, Pa sq density, kg m3

    r interfacial tension, Nm1

    Subscripts

    f fracture

    i initial

  • pore-network models of multiphase flow. Adv. Water Resour. 25,

    H.Sh. Behbahani et al. / Journal of Petroleum Science and Engineering 50 (2006) 21393810691089.

    Bourbiaux, B.J., Kalaydjian, F.J., 1990. Experimental study of cocur-

    rent and countercurrent flows in natural porous media. SPE

    Reserv. Eng. 5, 361368.

    Chimienti, M.E., Illiano, S.N., Najurieta, H.L., 1999. Influence of

    temperature and interfacial tension on spontaneous imbibition

    process. SPE 53668, Latin American and Caribbean Petroleum

    Engineering Conference, Caracas, Venezuela.

    Cil, M., Reis, J.C., Miller, M.A., Misra, D., 1998. An examination of

    countercurrent capillary imbibition recovery from single matrix-

    blocks and recovery predictions by analytical matrix/fracture

    transfer functions. SPE 49005, Annual Technical Conference,

    New Orleans, Louisiana, USA.

    Civan, F., 1994. A theoretically derived transfer function for oil recov-

    ery from fractured reservoirs by waterflooding. SPE 27745, Ninth

    Symposium on Improved Oil Recovery, Tulsa, Oklahoma, Aprilm matrix

    o oil

    r residual

    w water

    Acknowledgments

    We would like to thank the sponsors of the ITF

    project on dImproved Simulation of Flow in Fracturedand Faulted Reservoirs for funding this research. Has-

    san Behbahani would like to thank NIOC for providing

    financial support.

    References

    Al-Lawati, S., Saleh, S., 1996. Oil recovery in fractured oil reservoirs

    by low IFT imbibition process. SPE 36688, Annual Technical

    Conference and Exhibition, Denver, Colorado, USA.

    Aronofsky, J.S., Masse, L., Natanson, S.G., 1958. A model for the

    mechanism of oil recovery from the porous matrix due to water

    invasion in fractured reservoirs. Trans. AIME 213, 1719.

    Aziz, K., Settari, A., 1979. Petroleum Reservoir Simulation. Elsevier

    Applied Science, London.

    Babadagli, T., Ershaghi, I., 1992. Imbibition assisted two-phase flow

    in natural fractures. SPE 24044, Western Regional Meeting,

    Bakersfield, California, USA.

    Barenblatt, G.I., Zheltov, J.P., 1960. On the basic equations of single-

    phase flow of fluid through fractured porous media. Dokl. Akad.

    Nauk SSSR 132 (3), 545548.

    Barenblatt, G.I., Patzek, T.W., Silin, D.B., 2003. The mathematical

    model of non-equilibrium effects in wateroil displacement. SPEJ

    8 (4), 409420.

    Baldwin, B.A., Spinler, B.A., 2002. In situ saturation develop-

    ment during spontaneous imbibition. J. Pet. Sci. Eng. 35 (1-2),

    2332.

    Behbahani, H.Sh., Blunt, M.J., 2004. Analysis of imbibition in mixed-

    wet rocks using pore-scale modelling. SPE 90132, Annual Tech-

    nical Conference and Exhibition, Houston, Texas, USA.

    Blunt, M.J., Jackson, M.D., Piri, M., Valvatne, P.H., 2002. Detailed

    physics, predictive capabilities and macroscopic consequences for1720.Civan, F., Rasmussen, M.L., 2001. Asymptotic analytical solu-

    tions for imbibition waterfloods in fractured reservoirs. SPEJ

    6, 171181.

    Civan, F., Rasmussen, M.L., 2003. Modelling and validation of

    hindered-matrixfracture transfer for naturally fractured petroleum

    reservoirs. SPE 80918, Production and Operations Symposium,

    Oklahoma City, Oklahoma, USA.

    Cuiec, L., Bourbiaux, B.J., Kalaydjian, F.J., 1994. Oil recovery

    by imbibition in low-permeability chalk. SPE Form. Eval. 9,

    200208.

    Dake, L.T., 1978. Fundamentals of Reservoir Engineering. Elsevier,

    Amsterdam.

    de Swaan, A., 1978. Theory of waterflooding in fractured reservoirs.

    SPEJ, 117122 (April 1978).

    Di Donato, G., Blunt, M.J., 2004. Streamline-based dual-porosity

    simulation of reactive transport and flow in fractured reservoirs.

    Water Resour. Res. 40, W04203. doi:10.1029/2003WR002772.

    Di Donato, G., Huang, W., Blunt, M.J., 2003. Streamline-based dual

    porosity simulation of fractured reservoirs. SPE 84036, Annual

    Technical Conference, Denver, Colorado, USA.

    Du Prey, E.L., 1978. Gravity and capillarity effects on imbibition in

    porous media. SPEJ, 195206 (June 1978).

    Gupta, A., Civan, F., 1994. An improved model for laboratory mea-

    surement of matrix to fracture transfer function parameters in

    immiscible displacement. SPE 28929, Annual Technical Confer-

    ence, New Orleans, Louisiana, 2528 September.

    Hamon, G., Vidal, J., 1986. Scaling-up the capillary imbibition pro-

    cess from laboratory experiments on homogeneous and heteroge-

    neous samples. SPE 15852, European Petroleum Conference.

    London, UK.

    Huang, W., Di Donato, G., Blunt, M.J., 2004. Comparison of stream-

    line-based and grid-based dual porosity simulation. J. Pet. Sci.

    Eng. 43, 129137.

    Iffly, R., Rousselet, D.C., Vermeulen, J.L., 1972. Fundamental study

    of imbibition in fissured oil fields. SPE 4102, Annual Technical

    Conference, Dallas, Texas, USA.

    Jackson, M.D., Valvatne, P.H., Blunt, M.J., 2003. Prediction of wet-

    tability variation and its impact on flow using pore-to reservoir-

    scale simulations. J. Pet. Sci. Eng. 39, 231246.

    Kazemi, H., Gilman, J.R., Elsharkawy, A.M., 1992. Analytical

    and numerical solution of oil recovery from fractured reser-

    voirs with empirical transfer functions. SPE Reserv. Eng. 7,

    219227.

    Li, Y., Morrow, N.R., Ruth, D., 2003. Similarity solution for linear

    counter-current spontaneous imbibition. J. Pet. Sci. Eng. 39 (34),

    309326.

    Ma, S., Morrow, N.R., Zhang, X., 1997. Generalized scaling of

    spontaneous imbibition data for strongly water-wet systems. J.

    Pet. Sci. Eng. 18, 165178.

    Matejka, M.C., Llanos, E.M., Civan, F., 2002. Experimental determi-

    nation of matrix-to-fracture transfer functions for oil recovery by

    water imbibition. J. Pet. Sci. Eng. 33 (4), 253264.

    Mattax, C.C., Kyte, J.R., 1962. Imbibition oil recovery from fractured

    water drive reservoirs. SPEJ, 177184 (June 1962).

    Najurieta, H.L., Galacho, N., Chimienti, M.E., Illiano, S.N., 2001.

    Effects of temperature and interfacial tension in different produc-

    tion mechanisms. SPE 69398, Latin American and Caribbean

    Petroleum Engineering Conference, Buenos Aires, Argentina.

    Parsons, R.W., Chaney, P.R., 1966. Imbibition model studies on

    water-wet carbonate rocks. SPEJ, 2634 (March 1966).

    Penuela, G., Hughes, R.G., Civan, F., Wiggins, M.L., 2002. Time-dependent shape factors for secondary recovery in naturally frac-

  • tured reservoirs. SPE 75234, SPE/DOE Improved Oil Recovery

    Symposium, Tulsa, Oklahoma, USA.

    Pooladi-Darvish, M., Firoozabadi, A., 2000. Experiments and mod-

    elling of water injection in water-wet fractured porous media. J.

    Can. Pet. Technol. 39 (3), 3142.

    Rangel-German, E.R., Kovscek, A.R., 2002. Experimental and ana-

    lytical study of multidimensional imbibition in fractured porous

    media. J. Pet. Sci. Eng. 36, 4560.

    Rangel-German, E.R., Kovscek, A.R., 2003. Time-dependent ma-

    trixfracture shape factors for partially and completely immersed

    fractures. SPE 84411, Annual Technical Conference, Denver,

    Colorado, USA.

    Ruth, D., Morrow, N.R., Yu, L., Buckley, J., 2000. Simulation study

    of Spontaneous imbibition. International Society of Core Analysts

    Annual Meeting, Abu Dhabi, UAE, October.

    Ruth, D., Mason, G., Morrow, N.R., 2003. A numerical study of the

    influence of sample shape on spontaneous imbibition. Interna-

    tional Society of Core Analysts Annual Meeting, Pau, France,

    September.

    Tang, G.Q., Firoozabadi, A., 2001. Effect of pressure gradient and

    initial water saturation on water injection in water-wet and mixed-

    wet fractured porous media. SPE Reserv. Evalu. Eng. 4, 516524.

    Tavassoli, Z., Zimmerman, R.W., Blunt, M.J., 2005. Analytic analysis

    for oil recovery during counter-current imbibition in strongly

    water-wet systems. Porous Media Transp. 58, 173189.

    Terez, I.E., Firoozabadi, A., 1999. Water injection in water-wet

    fractured porous media: experiments and a new model with

    modified BuckleyLeverett Theory. SPEJ 4 (2), 135141.

    Valvatne, P.H., Blunt, M.J., 2004. Predictive pore-scale modelling of

    two-phase flow in mixed wet media. Water Resour. Res. 40,

    W07406. doi10.1029/2003WR002627.

    Viksund, B.G., Morrow, N.R., Ma, S., Wang, W., Graue, A., 1998.

    Initial water saturation and oil recovery from chalk and sandstone

    by spontaneous imbibition. SCA-9814, International Society of

    Core Analysts, The Hague, September.

    Warren, J.E., Root, P.J., 1963. The behavior of naturally fractured

    reservoirs. SPEJ, 245255. Trans. AIME 228 (September 1963).

    Zhang, X., Morrow, N.R., Ma, S., 1996. Experimental verification of

    a modified scaling group for spontaneous imbibition. SPE Reserv.

    Eng. 11, 280285.

    Zhou, X., Morrow, N.R., Ma, S., 2000. Interrelationship of wettabil-

    ity, initial water saturation, aging time, and oil recovery by

    spontaneous imbibition and waterflooding. SPEJ 5 (2), 199207.

    Zhou, D., Jia, L., Kamath, J., Kovscek, R.A., 2002. Scaling of

    counter-current imbibition processes in low permeability porous

    media. J. Pet. Sci. Eng. 33 (1-3), 6174.

    H.Sh. Behbahani et al. / Journal of Petroleum Science and Engineering 50 (2006) 2139 39

    Simulation of counter-current imbibition in water-wet fractured reservoirsIntroductionSimulation of 1D-counter-current imbibitionBase modelsExperimental data predicted by network modelling

    Simulation of 2D-counter-current imbibitionValidity of the correlations, Eqs. (1) and (5)Theory and simulation of fracture flow and imbibitionPrevious experimental studiesTheory for dual porosity systemsGrid-based simulation of fracture/matrix flowAnalytical and 1D numerical solutionsAnalysis of the results

    ConclusionsAcknowledgmentsReferences