throbber
Methodology to Predict the Initiation of
`Multiple Transverse Fractures from
`Horizontal Wellbores
`
`D.G. CROSBY, Z. YANG, S.S. RAHMAN
`University of New South Wales
`
`Abstract
`Multi-stage, transversely fractured horizontal wellbores have
`the potential to greatly increase production from low permeabili-
`ty formations. Such completions are, however, susceptible to
`problems associated with near-wellbore tortuosity, particularly
`multiple fracturing from the same perforated interval. A criteri-
`on, based on that by Drucker and Prager, has been derived,
`which predicts the wellbore pressures required to initiate sec-
`ondary multiple transverse hydraulic fractures in close proximity
`to primary fractures. Secondary fracture initiation pressures pre-
`dicted by this new criterion compare reasonably well with those
`measured during a series of unique laboratory-scale multiple
`hydraulic fracture interaction tests. Both the multiple fracture
`initiation criterion and the laboratory results suggest that close
`proximity of primary hydraulic fractures increases the initiation
`pressures of secondary multiple fractures by the order of only
`14%. This demonstrates that transversely fractured horizontal
`wellbores have limited capacities to resist the initiation of multi-
`ple fractures from adjacent perforations or intersecting hetero-
`geneities. Petroleum engineers can use the multiple fracture ini-
`tiation criterion when designing hydraulic fracture treatments to
`establish injection pressure limits, above which additional multi-
`ple fractures will initiate and propagate from the wellbore.
`
`Introduction
`
`A significant proportion of the worldwide recoverable hydro-
`carbon resource exists in reservoirs possessing permeabilities of
`less than one milli-Darcy (mD). At present, low production rates
`accompanying such poor permeabilities imply that, if hydrocar-
`bons are to be exploited economically, some form of permeability
`enhancement or stimulation must be carried out within these
`reservoirs. Even where initial permeabilities are relatively high,
`stimulation may still be required to overcome problems associated
`with localised permeability damage due to, for example, drilling
`mud invasion. Matrix acidisation and hydraulic fracturing remain
`the principal reservoir stimulation techniques.
`The advantages of horizontal wells in comparison with vertical
`wells have been extensively documented. Indeed, in an increasing
`number of fields throughout the world, the production of hydro-
`carbons is performed exclusively through horizontal wells. Whilst
`still a relatively rare form of completion, fractured horizontal
`wells are becoming more common in low permeability forma-
`tions. This is particularly so where surface geographies dictate
`that wells must deviate from central drill pads, such as in offshore
`
`FIGURE 1: Longitudinal fracture configuration.
`
`or arctic regions.
`Hydraulic fractures, regardless of their origin, always attempt
`to propagate in planes orthogonal to the minimum horizontal
`stress, in what is commonly referred to as the “preferred fracture
`plane.” However, while hydraulic fracture propagation planes are
`fixed, the horizontal wellbores from which they emanate may
`assume completely arbitrary orientations. Two limiting wellbore-
`fracture configurations are the focus of much attention:
`• “Longitudinal Fractures” propagate in planes parallel with
`wellbore axes, as illustrated in Figure 1. They form where
`horizontal wells are drilled parallel with the larger of the
`horizontal stresses (or parallel with the preferred fracture
`plane);
`• “Transverse Fractures” propagate in planes orthogonal to
`wellbore axes, as illustrated in Figure 2. They form where
`horizontal wells are drilled perpendicular to the larger of the
`horizontal stresses (or perpendicular to the preferred fracture
`plane).
`A number of studies have been carried out, comparing the pro-
`duction characteristics between fractured horizontal wells and
`fractured or unfractured vertical wells(1-5). In homogeneous reser-
`voirs, longitudinally fractured horizontal wells offer no apprecia-
`ble productive advantage over similarly fractured vertical wells.
`Only in thin, high permeability formations will longitudinally
`fractured horizontal wells significantly outperform fractured verti-
`cal wells(1).
`Alternatively, transversely fractured wells have the ability to
`greatly increase production rates by virtue of the fact that any
`number of fractures may be widely distributed along the length of
`horizontal wells, as illustrated in Figure 2, through multi-stage
`treatments. The reduced contact areas between horizontal well-
`
`PEER REVIEWED PAPER (“REVIEW AND PUBLICATION PROCESS” CAN BE FOUND ON OUR WEB SITE)
`
`68
`
`Journal of Canadian Petroleum Technology
`
`1 of 8
`
`Ex. 2063
`IPR2016-01517
`
`

`

`FIGURE 2: Transverse fracture configuration.
`
`bores and transverse fractures introduce additional “choke skin
`effects(2),” which hinders productivity and increases injection
`pressures. In general, at least three transverse hydraulic fractures
`are required to outproduce a single vertical well stimulated with
`similar dimensioned hydraulic fractures(6, 7).
`Unfortunately, transversely fractured horizontal wellbores are
`still plagued by a number of problems, most of which stem from
`the complex fracture geometries connecting the wellbore to the
`main fracture. These complex fracture geometries usually take the
`form of multiple fractures, twisted fractures, H- or S-shaped frac-
`tures(8, 9).
`The above complex fracture geometries are more commonly
`collectively known as “near wellbore tortuosity,” and result in
`narrower than anticipated fracture widths. Near-wellbore tortuosi-
`ty ultimately leads to unacceptably high fracture treatment pres-
`sures, proppant bridging and pre-mature near-wellbore screenout,
`shorter than expected final fracture lengths, and poor fracture con-
`ductivities. The origin of these fracture complexities may be
`traced back to the manner in which hydraulic fractures initiate
`from the wellbore. A considerable amount of research is currently
`devoted to understanding the dynamics of hydraulic fracture initi-
`ation. Indeed, many recently adopted field practices, such as well-
`bore break-down using highly viscous fluids(10), and decreased
`perforation densities(11), minimise fracture tortuosity by control-
`ling the fracture initiation process.
`A number of researchers(11-14), have suggested that every for-
`mation possesses an inherent “critical” or “threshold” fracture
`fluid injection pressure, above which secondary (or “auxiliary”)
`hydraulic fractures may initiate from natural fractures intersecting
`the main hydraulic fracture or wellbore. Formations with low
`threshold pressures may be susceptible to severe multiple fractur-
`ing and short overall hydraulic fracture lengths.
`The presence of such threshold pressures can be recognised as
`the flat sections on log-log plots of injection pressure vs. time.
`Nolte and Smith(12) suggested that flaws such as natural fractures
`behave as “pressure regulators,” which dilate and increase fracture
`fluid leak-off as fluid injection pressures approach threshold pres-
`sures. In addition, both Nolte & Smith(12) and Warpinski(14)
`derived criteria determining the conditions required to dilate nat-
`ural fractures intersecting main hydraulic fractures. Wells which
`have been hydraulically fractured in formations in the Wattenberg
`field of the Denver Basin in the United States display symptoms
`of natural fracture dilation with increasing fluid injection
`pressures(14).
`As a contribution to the above theories, this paper presents the
`results of analytical work carried out in order to establish the hori-
`zontal wellbore fluid pressures required to initiate additional,
`closely-spaced transverse multiple fractures from horizontal well-
`bores. In addition, the results of a series of unique laboratory mul-
`tiple fracture initiation tests are presented that provide support for
`the analytical work.
`
`Theoretical Considerations
`The stress state at the wall (in the direction of σL) of a pres-
`surised wellbore may be described in cylindrical coordinates by
`the following expressions:
`
`October 2001, Volume 40, No. 10
`
`FIGURE 3: Stresses at the wellbore wall.
`

`
`r
`
`wP=
`
`..................................................................................................(1)
`

`
`θ =
`

`3 L
`
`−
`

`
`−
`
`l
`
`wP
`

`
`zz
`
`=
`

`
`z
`
`(
`
`
`v2
`

`
`−
`
`−
`
`L
`

`
`l
`
`..............................................................................(2)
`)
`
`.......................................................................(3)
`
`where:
`σr
`= radial stress;
`σθ = tangential stress;
`σzz = wellbore axial stress;
`σL = larger in situ stress acting orthogonal to the axis of an
`arbitrarily oriented wellbore;
`σl
`= smaller in situ stress acting orthogonal to the axis of an
`arbitrarily oriented wellbore;
`σz
`= in situ stress acting along the axis of an arbitrarily orient-
`ed wellbore;
`Pw = wellbore fluid pressure;
`= Poisson’s ratio.
`v
`Compressive stresses are taken as positive. Figure 3 shows the
`orientation of the stresses described in Equations (1) to (3) on the
`wellbore wall. The classic Hubbert and Willis(15) expression,
`which makes use of Equation (2), has historically been used to
`estimate the wellbore fluid pressures required to initiate tensile
`hydraulic fractures from vertical wellbores:
`
`i
`Pwp
`
`≥
`
`3σ
`
`h
`
`−
`

`
`H
`
`+
`

`
`t
`
`...........................................................................(4)
`
`where:
`σH = maximum in situ horizontal stress;
`σh = minimum in situ horizontal stress;
`σt
`= rock tensile strength; and
`Pi
`wp = wellbore fluid pressure required to initiate a single
`(primary) fracture.
`This expression assumes that fractures form when the tangen-
`tial stress on the wellbore wall exceeds the rock tensile strength,
`and ignores the influence of pore pressure. Haimson and
`Fairhurst(16) derived a similar expression which accounted for
`poroelastic effects.
`Under the stress conditions usually prevailing at the wellbore
`wall during fracture initiation, the tangential and radial (equiva-
`lent to wellbore fluid pressure) stresses generally assume the roles
`of minor and major principal stresses respectively:
`

`
`r
`
`>
`

`
`zz
`
`>
`
`σθ
`
`........................................................................................(5)
`That is, the wellbore axial stress (σzz) is the intermediate princi-
`pal stress. The classic Hubbert and Willis expression [Equation
`
`2 of 8
`
`69
`
`Ex. 2063
`IPR2016-01517
`
`

`

`FIGURE 4: Stresses acting upon secondary fractures initiating in
`close proximity to a primary fracture.
`
`FIGURE 5: Distribution of stresses acting orthogonal to the plane
`of a radial, uniformly pressured fracture.
`
`As described previously by Equation (5), the radial (σr), axial
`(σzz) and tangential stresses (σθ) at wellbore walls assume the
`roles of major, intermediate and minor principal stresses, respec-
`tively, at depths typically of interest to the petroleum industry.
`Recall that, at the wellbore wall, radial stress is equivalent to well-
`bore fluid pressure. Therefore, in terms of the stresses acting on
`wellbore walls at failure, octahedral shear and normal stresses
`may be defined as follows:
`
`.........................(9)
`
`(
`
`i
`P
`wp
`
`−
`

`
`zz
`
`) +
`
`2
`
`(
`

`
`) +
`
`2
`
`(
`
`−
`

`

`
`zz
`

`

`
`−
`
`i
`P
`wp
`
`)
`
`2
`
`1 3
`

`
`oct
`
`=
`
`...................................................................(10)
`
`i
`P
`wp
`
`+
`

`
`+
`
`σθ
`
`zz
`
`)
`
`(
`
`1 3
`

`
`oct
`
`=
`
`where Pi
`wp is the wellbore fluid pressure required to initiate a sin-
`gle (primary) fracture. In addition, where the wellbore is oriented
`in a principal in situ stress direction, the tangential stress may be
`expressed in terms of the in situ stresses acting orthogonal to the
`wellbore axis, and the wellbore fluid pressure:
`

`
`θ =
`

`3 l
`
`−
`

`
`−
`
`iP
`wp
`
`L
`
`..........................................................................(11)
`
`By substituting Equations (9), (10) and (11) into the Drucker
`and Prager expression [Equation (6)], the following modified
`expression is derived:
`
`(
`
`) +
`
`2
`
`(
`

`
`−
`

`
`−
`
`3
`
`
`
`+
`
`) +
`
`2
`
`(
`

`
`3
`
`−
`

`
`−
`
`2
`
`L
`
`) =
`
`2
`
`(4)] ignores the influence of the intermediate principal stress.
`However, where secondary fractures are forced to initiate in the
`presence of a primary fracture (illustrated in Figure 4), the well-
`bore axial stress (intermediate principal stress) must play some
`role. Therefore, any criteria estimating the conditions leading to
`the initiation of closely-spaced secondary fractures must incorpo-
`rate intermediate principal stresses.
`The influence of the intermediate principal stress on rock fail-
`ure has been the subject of debate for a number of years.
`Experimental work carried out to establish the role of intermediate
`principal stresses on material failure requires relatively complex
`poly-axial test apparatus of the type used by Mogi(17). As the
`influence of the intermediate principal stress is expected to be
`subtle, highly accurate load measurement is required. This is in
`contrast with the more widely used and simpler tri-axial tests, in
`which the intermediate and minor principal stresses are assumed
`to be equal. In addition, the highly anisotropic nature of rock com-
`pounds the difficulties associated with understanding the role of
`intermediate principal stresses on rock failure. Despite the diffi-
`culties, experimental work, such as that performed by Mogi, sug-
`gests that increasing intermediate principal stresses (while main-
`taining constant minor principal stresses) increases the magni-
`tudes of major principal stresses at failure. Mogi also demonstrat-
`ed that the degree to which the intermediate principal stress influ-
`ences rock failure is closely linked to lithology.
`The Drucker and Prager(18) failure criterion (often referred to as
`the “Extended von Mises criterion”), which accommodates the
`influence of the intermediate principal stress, may be used as the
`basis of a secondary fracture initiation criterion. The Drucker and
`Prager failure criterion may be simply defined by the following
`expression:
`
`zz
`
`zz
`
`+σ σ
`l
`
`L
`
`i
`P
`wp
`
`l
`
`i
`P
`wp
`
`i
`P
`wp
`
`1 3
`
`+
`

`
`o
`
`(
`

`
`3
`
`l
`

`3
`
`−
`

`
`L
`
`+
`

`
`ZZ
`
`)
`
`...................................................................(12)
`
`The above expression may be solved for Pi
`wp.
`The wellbore axial stress (σzz) is a function of primary fracture
`proximity (2s) and pressure (Pfp). The local stresses exerted by
`radial fractures may be described analytically, such as by solu-
`tions derived by Sneddon(21), which is illustrated graphically in
`Figure 5. This figure demonstrates that localised stress increases
`induced by hydraulic fractures diminish rapidly with increasing
`distance away from the fractures. Therefore, unless secondary
`fractures initiate within close proximity to a primary fracture, the
`wellbore pressures required to initiate secondary fractures will be
`no greater than those required to initiate single, isolated hydraulic
`fractures. However, in many cases, particularly where the pres-
`surised intervals in horizontal wells are small, or during the
`hydraulic fracturing of vertical wells, secondary hydraulic frac-
`

`
`oct
`
`=
`
`τ ψσ+
`
`
`o
`
`oct
`
`..................................................................................(6)
`
`where:
`toct = “octahedral shear stress”
`
`(
`

`
`1
`
`−
`

`
`2
`
`) +
`
`2
`
`(
`

`
`−
`

`
`3
`
`2
`
`) +
`
`2
`
`(
`

`
`−
`

`
`3
`
`)
`
`2
`
`1
`
`1 3
`
`=
`
`...........................(7)
`σoct = “octahedral normal stress” (or “mean stress term”)
`
`+
`

`
`2
`
`+
`

`
`)
`
`3
`
`(
`

`
`1
`
`1 3
`
`=
`
`......................................................................(8)
`τo and ψ are material properties which must be determined experi-
`mentally. σ1, σ2 and σ3 are the major, intermediate and minor
`principal stresses, respectively. This criterion is widely used as a
`means of assessing wellbore stability(19, 20).
`
`70
`
`Journal of Canadian Petroleum Technology
`
`3 of 8
`
`Ex. 2063
`IPR2016-01517
`
`

`

`TABLE 1: Summary of the laboratory-scale fracture test parameter, and comparisons with field dimensions.
`
`Parameter
`
`Wellbore Dimensions:
`Wellbore radius, rw (m)
`
`Reservoir Properties:
`Permeability, K (mD)
`
`Geomechanical Properties:
`Elastic modulus, E (MPa)
`Poisson’s ratio, v
`Fracture toughness, KIc (
`
`MPa m
`
`)
`
`Laboratory-Derived Drucker & Prager Parameters:
`τo

`
`In situ Stresses:
`Vertical in situ stress, σv (MPa)
`Min. in situ horiz. Stress, σh (MPa)
`Max. in situ horiz. Stress, σH (MPa)
`
`Hydraulic Fracture Treatment Data:
`m/
`s
`)
`Leak-off, kl (
`Injection rate, Q (m3/s)
`Injection period, t (min)
`Fracture fluid viscosity, m (Pa s)
`
`Field-Scale Dimension
`
`Lab.-Scale Dimension
`
`0.091
`
`0.1
`
`50,000
`0.25
`3
`
`64
`47
`57
`
`4.8e-5
`0.072
`27
`0.5
`
`0.006
`
`8e-5
`
`700
`0.2
`0.27
`
`7.72
`0.73
`
`7
`5
`6
`
`3.8e-8
`3e-7 - 1e-6
`1
`30 (@ 23˚ C)
`
`Laboratory-Scale Fracture Studies
`Experimental Set-up
`A series of laboratory-scale hydraulic fracture experiments
`have been carried out in an attempt to model the initiation and
`propagation of, and interaction between, multiple transverse frac-
`tures. Only the fracture initiation aspects of the experimental work
`will be described in this paper. The fracture propagation issues
`will be discussed in future publications.
`In order to produce representative results, the laboratory tests
`adhered to the strict scaling procedures outlined by de Pater et
`al.(22) The laboratory test dimensions were scaled upon those typi-
`cally encountered during the exploitation of low permeability gas
`resources in Central Australia. Length dimensions were scaled
`according to wellbore radius (rw). Extremely low injection rates
`were employed in order to achieve stable fracture initiation and
`growth. This necessitated the use of an extremely high viscosity
`fracture fluid and a low leak-off fracture medium. Correct scaling
`also required that the fracture medium possess a low modulus (E)
`and toughness (KIC).
`A high viscosity (30 Pa•s at 23˚ C) industrial lubricant was
`employed as a fracture fluid. The fracture test material was a
`Portland cement-based material composed of the following:
`
`30%*
`Off-white Portland cement
`70%*
`Silica flour
`0.35*
`Water/cement ratio
`Acrylic emulsion/water ratio: 1:7.1*
`Water Reducing Agent:
`2,000 ml/100 kg Portland cement
`
`(* by weight of total cementicious material)
`The low cement content ensured that the fracture test material
`possessed a low elastic modulus (7,000 MPa) and low toughness
`(0.27
`). In addition, the high proportion of silica flour
`MPa m
`and inclusion of acrylic emulsion provided the test material with a
`negligible permeability (8e-5 mD). Table 1 summarises laboratory
`test material properties, and compares them with those in the field.
`
`4 of 8
`
`71
`
`Ex. 2063
`IPR2016-01517
`
`tures may indeed be forced to initiate in close proximity to prima-
`ry fractures. In addition, primary fractures initiated from wells in
`formations unbounded by more highly stressed intervals may
`grow to significant radii prior to secondary fracture initiation.
`Where secondary fractures initiate close to a primary fracture,
`the wellbore axial stress (σzz), against which the secondary frac-
`ture must form, approaches that of the primary fracture pressure.
`Under such ‘zero-spaced’ conditions, the single fracture Drucker
`and Prager criterion described by Equation (12) can be modified
`such that it assumes the following form:
`
`P
`
`fp
`
`−
`
`
`
`+σ σ
`l
`
`3
`
`L
`
`+
`
`i
`P
`ws
`
`) +
`
`2
`
`(
`

`
`3
`
`l
`
`−
`

`
`L
`
`−
`
`2
`
`i
`P
`ws
`
`) =
`
`2
`
`(
`
`−
`
`i
`P
`ws
`
`P
`
`fp
`
`) +
`
`2
`
`(
`
`1 3
`
`+
`

`
`o
`
`(
`

`3
`

`
`3
`
`l
`
`−
`

`
`L
`
`+
`
`fpP
`
`)
`
`....................................................................(13)
`
`where:
`Pfp = uniform primary fracture pressure;
`Pi
`ws = wellbore pressure required to initiate secondary hydraulic
`fracture.
`Equation (13) can then be solved for Pi
`ws. In the field, however,
`individual hydraulic multiple fractures are not hydraulically iso-
`lated. This is in contrast with the ‘static’ multiple fracture initia-
`tion described by Equation (13). Indeed, the fluid pressures within
`the primary and secondary fractures are coupled. Thus the modi-
`fied Drucker and Prager multiple fracture initiation criterion
`described by Equation (13) must be solved iteratively. The
`dynamic solution process simply involves substituting, at each
`iteration, the wellbore pressure required to initiate a secondary
`fracture (Pi
`ws), into the uniform pressure within the primary frac-
`ture (Pfp). This process is repeated until convergence.
`
`October 2001, Volume 40, No. 10
`
`

`

`pressure record for multiple fracture Test #3 is illustrated in
`Figure 9. The fracture pressures displayed in Figure 9 have been
`corrected for compressibility of the injection system and fracture
`fluid. During injection testing, both wellbores were pressurised at
`similar rates. A primary fracture eventually initiated from one of
`the two wellbores. Immediately after the primary fracture initiat-
`ed, the wellbore from which it emanated was shut-in. Injection
`into the primary fracture recommenced only after the secondary
`fracture (from the opposing wellbore) was initiated. Injection into
`both fractures was allowed to proceed at a constant rate for a peri-
`od of approximately two minutes, whereupon they were both
`shut-in.
`
`Results
`The Drucker and Prager material constants (τo and ψ) were
`derived through tri-axial testing of the fracture test material, and
`are listed in Table 1. By substituting these material constants and
`the applied in situ stresses (σl, σL and σz) into Equation (13), the
`following criterion is derived:
`(
`) +
`(
`) +
`
`−
`
`i
`P
`ws
`
`P
`
`fp
`
`2
`
`P
`
`fp
`
`−
`
`+
`
`11
`
`i
`P
`ws
`
`2
`
`−(
`
`11 2
`
`) =
`
`2
`
`i
`P
`ws
`
`1 3
`
`.+
`.
`7 72 0 245
`
`(
`
`)
`
`11
`
`+
`
`Pfp
`
`............................................................................(14)
`
`This criterion can be used to estimate the wellbore fluid pres-
`sure (Pi
`ws) required to initiate secondary hydraulic fractures in the
`presence of a nearby inflated primary fracture. The laboratory
`fracture test provided a unique opportunity to establish the validi-
`ty of this modified criterion.
`As described previously, field-scale primary and secondary
`fracture pressures are coupled. This contrasts with the experimen-
`tal configuration, in which the primary and secondary multiple
`fractures initiated and propagated in hydraulic isolation. However,
`
`FIGURE 6: Laboratory block configuration.
`
`The synthetic fracture test material was cast into blocks mea-
`suring 400 × 400 × 400 mm. Thin circular plastic disks were posi-
`tioned inside the blocks during casting (Figure 6). When intersect-
`ed by a wellbore, these disks constrained the location at which
`hydraulic fractures initiated. The disks were separated by a dis-
`tance 30 mm, and were intersected by two, hydraulically isolated
`wellbores, each possessing a radius of 6 mm. Stainless steel injec-
`tion tubes, of 3 mm internal radius, were grouted into the well-
`bores, leaving a small “open hole” section immediately adjacent
`to the plastic disks. The absence of a sand fraction left the synthet-
`ic fracture medium susceptible to shrinkage cracking. However,
`this was minimised by storing the blocks under conditions of
`100% humidity during curing, and until immediately prior to
`testing.
`
`The blocks were placed in a poly-axial cell, as illustrated in
`Figure 7. Through the use of water-filled flat-jacks, the poly-axial
`cell exerted stresses on all faces of each block. The magnitudes of
`the applied stresses (Table 1) were such that the wellbores were
`oriented in the direction of the minor principal stress. This orien-
`tation between wellbore and stresses promoted the formation of
`transverse hydraulic fractures. Two computer-controlled linear
`displacement pumps, shown in schematic form in Figure 8, were
`used to inject fracture fluid independently into each wellbore. In
`addition, transducers independently measured the initiation and
`propagation pressures of each multiple fracture. The injection
`
`FIGURE 7: Poly-axial cell configuration.
`
`FIGURE 8: Schematic of the laboratory hydraulic fracture
`injection and data recording systems.
`
`72
`
`Journal of Canadian Petroleum Technology
`
`5 of 8
`
`Ex. 2063
`IPR2016-01517
`
`

`

`FIGURE 9: Injection pressure record for the laboratory multiple fracture initiation and propagation test #3.
`
`ple fractures from adjacent perforations or intersecting hetero-
`geneities. That is, second, third, and fourth, etc., generation
`hydraulic transverse fractures will be generated, in turn, as the
`wellbore pressure reaches the required threshold pressure (Pi
`ws).
`This study contributes to previous works by establishing the
`wellbore pressures at which additional multiple fractures will ini-
`tiate from an initially intact wellbore. Such information may be
`invaluable where transversely fractured horizontal wells are to be
`completed in naturally fractured formations. It is emphasized,
`however, that this technique is yet to be trialled in the field. Such
`a trial may be performed on a single horizontal wellbore, and will
`require measurement of Drucker and Prager material constants of
`the formation and a means of detecting or inferring the presence
`of multiple hydraulic fractures, through net treating pressure
`analysis.
`A hypothetical fracture injection pressure profile is shown in
`Figure 10, based on the average initiation and propagation pres-
`sures of transversely fractured horizontal wells in the North Sea(9).
`The measured average fracture initiation pressure was of the order
`of 46.54 mPa (6,750 psi). Therefore, according to the modified
`Drucker and Prager criterion [Equation (13)], it is estimated that
`secondary multiple fractures will initiate if the bottom hole treat-
`ing pressure is allowed to exceed 53.06 mPa (7,695 psi). Such
`
`due to the modest impact of the intermediate principal stress on
`fracture initiation pressures, the difference between the iterative
`and static solutions of Equation (14) should not be unduly large.
`Thus Equation (14) was solved in a static fashion to arrive at a
`secondary fracture initiation pressure (Pi
`ws) of 22.4 MPa. The theo-
`retical Drucker and Prager primary (or single) fracture initiation
`pressure (Pi
`wp) for the laboratory test may be easily derived
`through use of Equation (12), though with the wellbore axial
`stress (σzz) estimated through use of Equation (3). Accordingly,
`the analytically derived ratio of secondary to primary fracture
`
`for the DC-3 laboratory configuration
`
` 
`
`i w
`P P
` 
`
`ws
`i
`
`p
`
`initiation pressure,
`was 1.14.
`
`The experimentally-derived ratio of secondary to primary frac-
`ture initiation pressure for Test DC-3 was approximately 1.08.
`The reason for the slight disparity between the analytically deter-
`mined (1.14) and experimentally measured (1.08) secondary to
`primary fracture initiation pressure ratios is most likely due to
`some incorrect assumptions regarding the experimental work. The
`analytical Drucker and Prager solution [Equation (14)] assumes
`that the primary fracture is uniformly pressurised to a degree
`equal to its break-down pressure (20.45 MPa). However, perusal
`of the experimental pressure plot (Figure 13) indicates that upon
`shut-in, the primary fracture pressure decreased significantly. By
`the time the secondary fracture initiated, the primary fracture pres-
`sure had decreased from 20.45 MPa to 14.9 MPa. Therefore, if Pfp
`= 14.9 MPa is substituted into Equation (14), the analytically
`
`i w
`P P
`
`ws
`i
`
`derived ratio of
`
`equals 1.11, which compares more closely
`
`p
`with the experimentally determined ratio of 1.08.
`It is important to note that this criterion deals with fracture ini-
`tiation, and as such makes no comment regarding the eventual
`size of secondary hydraulic fractures in comparison with the pri-
`mary fracture. The exact size and distribution of multiple
`hydraulic fractures is still the source of constant debate, and was
`beyond the scope of this study.
`
`Discussion
`The reasonable agreement between the analytical estimates and
`experimental measurements of secondary fracture initiation pres-
`sure provides confidence in the validity of the modified Drucker
`and Prager criterion [Equation (14)]. Both the multiple fracture
`initiation criterion and the laboratory results suggest that close
`proximity of primary hydraulic fractures increases the initiation
`pressures of secondary multiple fractures by the order of only
`14%. This demonstrates that transversely fractured horizontal
`wellbores have limited capacities to resist the initiation of multi-
`
`October 2001, Volume 40, No. 10
`
`FIGURE 10: Hypothetical injection pressure record, illustrating of
`the onset of secondary fracture initiation with increasing wellbore
`pressure. Primary fracture initiation and propagation magnitudes
`are based on field data obtained during the treatment of
`transversely fractured horizontal wells in the North Sea(9).
`Secondary fracture initiation pressures were estimated through
`use of the modified Drucker and Prager criterion [Equation (13)].
`
`6 of 8
`
`73
`
`Ex. 2063
`IPR2016-01517
`
`

`

`high fracture propagation pressures will occur if fractures are
`propagated through highly confined formations or in the presence
`of perforation or fracture tortuosity flow restrictions. Field evi-
`dence(8) suggests that the formation of multiple hydraulic fractures
`from such completions is particularly detrimental to successful
`fracture treatments and post-fracture productions. Fracture design
`engineers can use secondary fracture initiation pressures furnished
`by the modified Drucker and Prager criterion to design fracture
`treatments that maintain fracture and wellbore pressures below
`that required to initiate additional multiple hydraulic fractures.
`Such limiting injection pressures will be a complex function of in
`situ stress distribution, reservoir thickness and rock properties.
`Modified fracture designs may incorporate reduced fluid injection
`rates or fracture fluid viscosities, which can result in final
`hydraulic fracture lengths which are significantly smaller than
`those originally desired. The alternative to limiting the generation
`of mulitple secondary fractures is the increased risk of premature
`screen-out, or time-consuming and costly proppant slug exercises.
`
`Conclusions
`1. A criterion, based on that by Drucker and Prager, has been
`derived, which predicts the wellbore pressures required to
`initiate secondary multiple transverse hydraulic fractures in
`close proximity to primary fractures.
`2. Secondary fracture initiation pressures predicted by this new
`criterion compares reasonably well with those measured dur-
`ing laboratory-scale multiple hydraulic fracture interaction
`tests.
`3. Both the multiple fracture initiation criterion and the labora-
`tory results suggest that close proximity of primary
`hydraulic fractures increases the initiation pressures of sec-
`ondary multiple fractures by the order of only 14%. This
`demonstrates that transversely fractured horizontal wellbores
`have limited capacities to resist the initiation of multiple
`fractures from adjacent perforations or intersecting
`heterogeneities.
`4. Designers of hydraulic fracture treatments can use the multi-
`ple fracture initiation criterion to establish injection pressure
`limits (or “threshold pressures”), above which additional
`multiple fractures will initiate and propagate from the well-
`bore. Such a tool may be particularly useful in the design
`and treatment of transverse fractures from horizontal well-
`bores, which are particularly prone to complications associ-
`ated with multiple fracturing.
`
`Acknowledgements
`The authors wish to acknowledge the Australian Petroleum
`Cooperative Research Centre (APCRC) for supporting this work
`and CSIRO Petroleum for use of their laboratory facilities in
`Melbourne.
`
`NOMENCLATURE
`= wellbore radius;
`rw
`s
`= half-distance between primary fracture and site of
`secondary fracture initiation;

`= Poisson’s ratio;
`Pfp = primary fracture pressure;
`Pw = wellbore fluid pressure;
`Pi
`wp = wellbore pressure required to initiate a single (primary)
`hydraulic fracture;
`Pi
`ws = wellbore pressure required to initiate secondary hydraulic
`fracture;
`= hydraulic fracture radius;
`Rf
`σ1 = major principal stress;σ
`σ2 = intermediate principal stress;
`σ3 = minor principal stress;
`σv = vertical in situ stress;
`
`σH = maximum horizontal in situ stress;
`σh = minimum horizontal in situ stress;
`σz
`= in situ stress acting along the axis of an arbitrarily
`oriented wellbore;
`σzz = axial wellbore stress;
`σl
`= smaller in situ stress acting orthogonal to the axis of an
`arbitrarily oriented wellbore;
`σL = larger in situ stress acting orthogonal to the axis of an
`arbitrarily oriented wellbore;
`σr
`= radial stress;
`σt
`= rock tensile strength;
`σθ = tangential stress;
`τoct = octahedral shear stress;
`σoct = octahedral normal stress;
`τo
`= Drucker and Prager material constant;
`ψ = Drucker and Prager material constant;
`
`SI Metric Conversion Factors
`ft × 3.048*
`E + 00 = m
`in × 2.54*
`E + 00

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket