5
THE LEADING EDGE September 2007 Vol. 26, No. 9 Pages 91081-1232 Fractures September 2007, Vol. 26, No. 9 Special Section: Special Section: Fractures THE SOCIETY OF EXPLORATION GEOPHYSICISTS The international society of applied geophysics ISSN 1070-485X 2609_c1_cover1.qxd 10/9/07 10:56 AM Page 1

Special Section: Fractures - Purdue University...T HE L EADING E DGE September 2007 Vol. 26, No. 9 Pages 91081-1232 Fractures September 2007, Vol. 26, No. 9 Special Section: Fractures

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

  • THE

    LEADING

    EDGE

    September 2007 Vol. 26, N

    o. 9Pages 91081-1232

    Fractures

    September 2007, Vol. 26, No. 9

    Special Section: Special Section:

    Fractures

    THE SOCIETY OF EXPLORATION GEOPHYSICISTSThe international society of applied geophysicsISSN 1070-485X

    2609_c1_cover1.qxd 10/9/07 10:56 AM Page 1

  • F ractures in rock cause seismicanisotropy because they are orientedmechanical discontinuities that mayoccur in sets with preferred orienta-tions. However, anisotropy from ori-entation can be altered or masked bygradients in stress and fluids. Suchgradients in stress occur naturally inthe Earth or are induced throughanthropogenic activities. For instance,stress increases with depth in the Earthbecause of lithostatic forces, and in agiven region, local gradients in hori-zontal stresses can arise through thetectonic history of the region (e.g.,folding, faulting, etc.). In addition, gradients in fluid dis-tributions as well as pore pressures occur when fluids areeither injected or withdrawn from a rock formation.

    Gradients in stress are particularly significant for frac-tures because fractures are weakly coupled through thepoints of contact between the two fracture surfaces. Thismeans that small physical modifications of the fracture cancause large changes in the mechanical, seismic, and hydraulicresponse of a fracture system. For instance, even lowamounts of fluid saturation at the level of only 10% can dou-ble seismic transmission across a fracture. This occursbecause the fluid provides a transmission medium aroundexisting asperities, couple asperities, or it couples asperitiesacross the fracture aperture. For such a weak-coupling sys-tem, the fracture geometry plays an essential role. The geom-etry defines not only the seismic properties of the fracture,but also the fluid flow through the fracture, which can behighly sensitive to stress. For example, the flow pathsthrough the fracture are composed of the apertures and spa-tial distribution of regions of contacts that are sensitive tochanges in stress. A factor of two decrease in aperture froman increase in stress on a fracture can lead to an order ofmagnitude decrease in the volumetric flow rate.

    To understand these stress-induced changes on the frac-ture geometry, a useful quantity is fracture-specific stiff-ness. Stiffness relates the deformation of a fracture (closingapertures and increasing asperity contacts) to the stress.Stiff fractures displace less, while compliant fractures dis-place more. Theoretically, a fracture can be represented bya displacement discontinuity, i.e., stresses across the frac-ture are continuous but the displacements are not. The dis-continuity in displacement is inversely proportional to thestiffness of the fracture. From experimental and numericalstudies it has been shown that there is a general correlationof fracture geometry and fracture stiffness: compliant frac-tures tend to have larger apertures and fewer or smallerregions of contact than fractures that are stiffer. It is impor-tant to keep in mind that this correlation is statistical andnot deterministic. Specific counter-examples certainly existthat show the opposite trend, but these special cases do notalter the average connection between fracture geometry andfracture stiffness.

    The displacement-discontinuity theory has been used toderive plane-wave transmission and reflection coefficientsfor fractured media, group time delays, compressional-modeand Rayleigh-mode interface waves, guided Love waves,and dispersion relationships. Many of these studies haveassumed that fracture stiffness is uniform along a fractureand uniform among fractures within a set. But, as describedearlier, because stress gradients commonly exist or areinduced in the Earth, one must consider what effects stressgradients have on seismic wave propagation in fracturedmedia. Gradients in stress cause gradients in fracture-spe-cific stiffness and ultimately affect any interpretation of seis-mic anisotropy.

    If fracture-specific stiffness varies along a fracture, thetime delay of a seismic wave also varies along the fractureplane. For example, consider a fracture with a linearly vary-ing fracture-specific stiffness (Figure 1a) caused by a gradi-ent in stress on the fracture. A plane wave propagatingupward from below is refracted because the time-delay isa function of fracture-specific stiffness. A low fracture stiff-ness produces a larger time delay than a region of the frac-ture with high stiffness. While this refraction certainlyviolates Snell’s law, it does not violate Fermat’s principle ofthe path of least time.

    An interesting example of the effect of stress gradienton seismic wave propagation across a fracture was studiedby Oliger et al. (2003). They showed that a single plane frac-ture with an axially symmetric stress distribution (radial gra-dient in stress) behaves as a seismic lens that focuses seismicenergy to a beam ‘‘waist’’ at a focal plane. A schematic ofthis concept is shown in Figure 1b. The fracture in Figure1b has a stiffness that increases quadratically away from thecenter. The low stiffness in the center of the fracture delaysthe waves relative to the stiffness farther out along the frac-ture plane. For a wave propagating upwards from below,the wave is focused by the fracture. The location of thereceiver in front of, at, or behind the focal plane determineswhether a converging, planar, or diverging wave front isobserved. Their work demonstrated that a 2D planar frac-ture with a stress gradient (contrasted with three-dimen-sional geologic structures such as basins and domes) canfocus seismic waves. Focusing of seismic waves by fractures

    Fracture anisotropy: The role of fracture-stiffness gradientsLAURA J. PYRAK-NOLTE, Purdue University, West Lafayette, Indiana, USA

    1124 THE LEADING EDGE SEPTEMBER 2007

    Figure 1. A fracture with (a) a linearly varying fracture-specific stiffness (high on the left to lowon the right) and (b) a radially varying fracture-specific stiffness (low in the center and increasesquadratically radially outward).

  • should be considered in the interpretation of seismic datafrom fractured strata with heterogeneous stress distributions.

    Another example of the effect of stress gradients on theinterpretation of seismic data from a set of parallel fracturescan be found in Hildyard and Young (2002). They showedthrough a comparison of data and numerical modeling thatknowledge of the state of stress in fractured systems is keyfor accurate interpretation of seismic data. In an attempt toreproduce, numerically, wave attenuation for shear andcompressional waves propagated parallel or perpendicularto a set of parallel fractures, they found that stress distrib-

    ution was of critical importance. Based on an analysis of thestress distribution in the experiment, the fracture-specificstiffness could not be assumed to be either the same amongthe fractures in the set, nor uniform along any single frac-ture within the set (Figure 2). In their analysis, the stress var-ied from 9 MPa to 30 MPa across the fracture set, and alongthe central fracture the stress varied from 3 MPa near theedges to 9 MPa in the center. By using a stress-dependentfracture-specific stiffness, they were able to match the lab-oratory data.

    While stress gradients can be eliminated or created inthe laboratory through experimental design, stress distrib-utions in the subsurface are set by lithostatic pressures,changes in pore pressure, as well as the local tectonic set-ting. Because wave attenuation and signal delay stronglydepend on the stiffness of the fracture, gradients in stresscertainly have significant consequences for the seismic inter-pretation of fracture properties in fractured reservoirs wherethe local stress field is perturbed during production.

    Another strong fracture anisotropy arises when seismicwaves are propagated parallel to fracture sets. Even whenthe fracture spacing is smaller than a wavelength, strongwaveguiding can occur when the fracture-specific stiffnessvaries along the set. Figure 3 shows a set of parallel frac-tures with a fracture spacing much smaller than a wave-length where the fracture stiffness is high in the centralfractures but decreases with distance from the center. Thisspatial variation in fracture-specific stiffness can producewaveguiding (as shown in Figure 3). Because of the gradi-ents in stiffness, wave transmission across the fracturesdecreases and the waves become internally reflected withina group of fractures producing a wave guide. If the gradi-ent in stiffness were similar to what is known in optics as aGRIN (graded index), the wave could snake though the lay-ers.

    SEPTEMBER 2007 THE LEADING EDGE 1125

    Figure 2. The stress distribution (for stress component normal to thefracture plane) in a set of parallel fractures under bi-axial loading(after Hildyard and Young, 2002).

    Figure 3. A gradient in fracture stiffness (red=high stiffness decreasingto black=low fracture stiffness) among fractures in a set where thefracture spacing is much smaller than a wavelength.

    Figure 4. Images of the acoustic wavefront propagated through (a) anintact sample F0, (b) fracture sample F3 in the dry condition, and (c)fracture sample F3 in the saturated condition. The fractures in theimages for sample F3 are oriented parallel to the time axis and arespaced 3 mm apart (i.e., approximately 20 fractures occur between 0and 60 mm). All of the wavefront images have a common time axis.The color scale on the right of each image represents the amplitude involts. Note that the color scale is different for images (a), (b), and (c).

  • In a fracture system, energy confinement that spans sev-eral fractures is possible either with a stress-induced distri-bution of fracture-specific stiffness as shown in Figure 3 orwith an uneven fluid saturation of fractures within a sys-tem of fractures. Water saturation increases both the nor-mal and shear fracture-specific stiffnesses. As describedearlier, a fracture can be viewed as being weakly coupledthrough the points of contact between the two surfaces. Theaddition of a fluid to this weakly coupled system signifi-cantly enhances the fracture stiffness. For normal stiffness,the coupling is enhanced through the bulk modulus of thefluid filling the fracture (e.g., water relative to air). For shearstiffness, the amount of enhancement is a function of thegeometry of the fracture. For example, if the fracture sur-faces are rough and interlocking, the void shapes mightenhance shear-wave transmission (i.e., like a wet clutch). Forsmooth parallel surfaces, little enhancement in shear-wavetransmission would be expected because shear waves decayin fluids. Thus, if a set of fractures were not uniformly sat-urated, a variation of both normal and shear fracture-spe-cific stiffness can occur.

    An example of waveguiding induced in a set of paral-lel fractures caused by uneven saturation within the set offractures is given by Xian et al. (2001). They used wavefrontimaging on aluminum samples with multiple parallel syn-thetic fractures to examine the effect of stress and satura-

    tion on wave propagation in amedium containing a set of par-allel fractures. Aluminum sam-ples were used, so all effectsobserved are the result of thefactures and not of the back-ground medium. Figure 4shows the acoustic wavefrontrecorded for an intact referencesample, F0, and a fracture sam-ple, F3. The fracture sample wascomposed of 28 fractures with afracture spacing of 3 mm, a spac-ing that was roughly equal to aquarter of a wavelength at a fre-quency of 0.5 MHz. The intactsample exhibited a uniformwavefront (Figure 4a) becauseof the isotropy of the aluminum.On the other hand, for the dryfracture sample in Figure 4b, thefirst-arriving wavefront had asignificantly smaller amplitudethan the intact sample, and thewaveform was mostly confinedto within the central region span-ning several fractures. There isa systematic trend in confine-ment as a function of frequency:the high-frequency componentsof the signal were the moststrongly confined, while thedominant low-frequency energywas more weakly confined andhad an earlier first arrival thanthe high-frequency componentsof the signal.

    An anomalous behavioroccurred when the fractureswere exposed to water with the

    purpose of saturating the fractures. Because of trapped airin the fractures, the water saturation was incomplete.Furthermore, there were gradients in the fracture saturationacross the set, which enhanced wave confinement rather thansuppressing it. In the saturated condition (Figure 4c), thewavefront for the fracture sample was more strongly con-fined than the wavefront in the dry condition and was muchlarger in amplitude. This apparently anomalous behaviorarose because the dominant low-frequency energy was con-fined by two fractures distant from the center of the sam-ple, approximately at locations of 20 mm and 40 mm. Thecentral fractures exhibited a larger value of stiffness thanthe fractures farther from the center, which confined thewave. These data illustrate that stiffness gradients in a setof parallel fractures can induce waveguiding even when thefracture spacing is much smaller than a wavelength.Interpretation of seismic data from fractured media wouldrequire both a knowledge of the local stress distribution aswell as possible gradients in saturation.

    Certainly, the opposite effect is more usually the case, inwhich saturation with water would be expected to decreaseconfinement anisotropy. As an example of this, Xian (2001)observed that a combination of stress and saturation couldmask the presence of fractures. He used an aluminum frac-ture sample that had seven fractures with a fracture spacingequal to a wavelength (12 mm for a frequency of 0.5 MHz).

    1126 THE LEADING EDGE SEPTEMBER 2007

    Figure 5. Acoustic wave front images are shown for fracture sample F12: (a) dry at 1.4 MPa, (b) dry at7.0 MPa, (c) water-saturated at 1.4 MPa, and (d) water-saturated 7.0 MPa. The images on the left showthe spatial distribution of energy for each condition taken at 63.5 microseconds. The images on the rightrepresent a 20-microsecond window of the wavefront as a function of horizontal position (taken at avertical position of 30 mm in the images on the left). The color scale associated with each image repre-sents the amplitude in volts. Please note that the color scale is different for each image.

  • In Figure 5 (left), the fractures were oriented parallel to thevertical axis and were located at horizontal positions of approx-imately 6 mm, 18 mm, 30 mm, 42 mm, and 54 mm, and sim-ilarly for the right side of the figure.

    The wavefronts for the dry fracture sample comparedat low and high stresses (Figures 5a and b) show that increas-ing pressure on the sample decreased the confinementanisotropy. At low stress, the acoustic wavefront in the frac-ture sample was strongly confined to the central layer. Forthis sample, the source was located at about 26 mm on thehorizontal axis and about 30 mm on the vertical axis. Theacoustic wavefront was delayed by the fractures and spreadout faster within the central confining layer than across thefractures. Energy confinement within the central layer wasalso observed in the image of the propagating wavefront(Figure 5a on the right). At this low stress, the low fracture-specific stiffness prevents any significant energy transmis-sion across the fractures. The arrival time of the compressionalmode wavefront is approximately the same as for the intactsample (Figure 4a) while the wavefronts outside the centralwaveguide are delayed and attenuated by each fracture thatthe wavefront crosses.

    On the other hand, at high stress (Figure 5b), theincreased fracture-specific stiffness allows greater trans-mission of energy across the fractures, less delay of thewavefront, and less energy confinement in the central layerbetween fractures. The nonsymmetry in the amplitude inthe image of the acoustic wavefront (Figures 5a and b) alsoindicates that the fracture-specific stiffness for each fracture,and even for different locations on the same fracture, werenot equal. For example, Figure 5b shows that the fractureat a horizontal position of 30 mm had a different stiffnessthan the fracture at a horizontal position of 18 mm basedon the asymmetry of the wavefront. Thus, symmetric posi-tions in the wavefront relative to the central layer do nothave the same energy.

    Homogeneous saturation of the fractures with water, com-bined with high stress, removes the strong effect of fractureson the acoustic wavefront and makes the fractures nearlyinvisible. By comparing Figures 5a and 5c (or Figures 5b and5d) for the same confining pressure, more energy is propa-gated across the saturated fractures than across the dry frac-tures. Increasing the stiffness of a fracture further increasesthe transmission of energy across the fracture and reduces theamount of energy that is internally reflected into a guidedmode. At a confining pressure of 7.0 MPa for the saturatedcondition (Figure 5d), the first-arriving wavefront is nearly uni-form and the presence of fractures is only observed in the laterarrivals. The wavefront exhibited a slight ellipticity (Figure5d on the left), although the high stiffnesses of the fracturesenabled the wavefront to spread out almost as if there wereno fractures. The energy distribution was observed to beroughly symmetric and only a slight delay in the wavefrontoccurred as it propagated across the fractures. This demon-strates that seismic anisotropy caused by parallel sets of frac-tures can be masked by saturating the fractures with a liquid.Hence, application of both stress and fluid saturation (homo-geneously) essentially erases the effects of the fractures on apropagating wavefront.

    In summary, I have shown that the ability to interpretfracture properties from seismic data is linked to spatial vari-ations in fracture-specific stiffness. Several examples of theeffect of gradients in fracture-specific stiffness on seismicwave propagation were presented to convey a sense that gra-dients in fracture-specific stiffness play an important rolein seismic anisotropy caused by fractures. Fracture-specificstiffness is intimately linked to the many different length

    scales associated with fracture geometry (apertures, spatialcorrelations, contact area, etc.) and to how these lengthscales are altered through physical processes. The lengthscales associated with the examples given in this paper aremuch smaller than those encountered in the field. However,an understanding of length scales associated with fracturesand fracture sets relative to seismic length scales is relevantbecause it determines when a fractured medium can betreated as an effective medium, or when a discrete fractureapproach is necessary. In addition, it is important to recog-nize that physical processes (such as stress, fluids, etc.) caneither homogenize a fracture stiffness, or it can induce greatercomplexity within a fracture or among fractures within aset. Therefore, a system that initially may be represented asan effective medium may fail to be so as alterations to a frac-tured reservoir proceed, and vice versa. A fundamentalunderstanding of how gradients in fracture-specific stiffnessalter the seismic response of a fractured medium will helpimprove the interpretation of seismic data and to recognizedeviations from homogeneity.

    Suggested reading. “Modeling seismic waves around under-ground openings in fractured rock” by Hildyard and Young(Pure and Applied Geophysics, 2002). “Focusing of seismic wavesby a single fracture” by Oliger et al. (Geophysical Research Letters,2003). “Transmission of seismic waves across single naturalfractures” by Pyrak-Nolte et al. (Journal of Geophysical Research,1990). “Single fractures under normal stress: The relationbetween fracture-specific stiffness and fluid flow” by Pyrak-Nolte and Morris (International Journal of Rock Mechanics andMining Sciences, 2000). “Compressional waves guided betweenparallel fractures” by Xian et al. (International Journal of RockMechanics and Mining Sciences, 2001). “Wavefront imaging ofenergy confinement by multiple parallel fractures” by Xian(MS thesis, Purdue University, 2001). TLE

    Acknowledgments: The author acknowledges support of this work by theGeosciences Research Program, Office of Basic Energy Sciences, U.S.Department of Energy (DE-FG02-97ER14785 08).

    Corresponding author: [email protected]

    SEPTEMBER 2007 THE LEADING EDGE 1127