`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`Copyright
`
`by
`
`Somnath Mondal
`
`2010
`
`
`
`
`
`IWS EXHIBIT 1060
`
`EX_1060_001
`
`
`
` The Thesis Committee for Somnath Mondal
`
` Certifies that this is the approved version of the following thesis:
`
`
`
`
`
` Pressure Transients in Wellbores: Water Hammer Effects and
`
`Implications for Fracture Diagnostics
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
` APPROVED BY
`
` SUPERVISING COMMITTEE:
`
`
`
`
`
`Supervisor:
`
`
`Mukul M. Sharma
`
`David DiCarlo
`
`
`
`IWS EXHIBIT 1060
`
`EX_1060_002
`
`
`
`
`
` Pressure Transients in Wellbores: Water Hammer Effects and
`
`Implications for Fracture Diagnostics
`
`
`
`
`
`by
`
` Somnath Mondal, B.E.
`
`
`
`
`
` Thesis
`
`Presented to the Faculty of the Graduate School of
`
`The University of Texas at Austin
`
`in Partial Fulfillment
`
`of the Requirements
`
`for the Degree of
`
`
`
` Master of Science in Engineering
`
`
`
`
`
` The University of Texas at Austin
`
` December, 2010
`
`IWS EXHIBIT 1060
`
`EX_1060_003
`
`
`
`Dedication
`
`
`
`To my family and friends.
`
`
`
`IWS EXHIBIT 1060
`
`EX_1060_004
`
`
`
`Acknowledgements
`
`
`
`It is a pleasure to thank those who made this thesis possible. I owe my deepest
`
`gratitude to my supervisor, Dr. Mukul M. Sharma, for his constant encouragement,
`
`guidance and support over the past two years. I would like to thank Dr. David DiCarlo for
`
`taking time to read and review this work.
`
`I would like to thank the members of the Hydraulic Fracturing and Sand Control
`
`Joint Industry Project (JIP) at the University of Texas for their financial support. I would
`
`like to specifically thank Mr. Adi Venkitaraman of Chevron for providing field data. I
`
`also express my sincere gratitude to Dr. George Wong of Shell for his expertise that has
`
`been immensely helpful in formulating this problem, and also for providing field data.
`
`I would like to acknowledge Jin Lee for her support in maintaining this wonderful
`
`research group. I thank my family and friends for making me who I am.
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`December, 2010
`
`
`
`v
`
`IWS EXHIBIT 1060
`
`EX_1060_005
`
`
`
`Abstract
`
`
`
` Pressure Transients in Wellbores: Water Hammer Effects and
`
`Implications for Fracture Diagnostics
`
`
`
`
`
`
`
`
`
`Somnath Mondal, MSE
`
`The University of Texas at Austin, 2010
`
`
`
`Supervisor: Mukul M. Sharma
`
`
`
`A pressure transient is generated when a sudden change in injection rate occurs
`
`due to a valve closure or injector shutdown. This pressure transient, referred to as a water
`
`hammer, travels down the wellbore, is reflected back and induces a series of pressure
`
`pulses on the sand face. This study presents a semi-analytical model to simulate the
`
`magnitude, frequency and duration of water hammer in wellbores. An impedance model
`
`has been suggested that can describe the interface, between the wellbore and the
`
`formation. Pressure transients measured in five wells in an offshore field are history
`
`matched to validate the model. It is shown that the amplitude of the pressure waves may
`
`be up to an order of magnitude smaller at the sand face when compared with surface
`
`measurements. Finally, a model has been proposed to estimate fracture dimensions from
`
`water hammer data.
`
`
`
`vi
`
`IWS EXHIBIT 1060
`
`EX_1060_006
`
`
`
`Table of Contents
`
`List of Tables ......................................................................................................... ix
`
`List of Figures ..........................................................................................................x
`
`Chapter 1: Introduction ............................................................................................1
`
`1.1 Literature Review......................................................................................5
`
`1.1.1 Water hammer Modeling ..............................................................5
`
`1.1.2 Fracture Impedance .......................................................................8
`
`Chapter 2: Model Formulation...............................................................................10
`
`2.1 Water hammer Modeling Equations .......................................................10
`
`2.1.1 Continuity Equation ....................................................................10
`
`2.1.2 Equation of Motion .....................................................................11
`
`2.1.3 Velocity of Water Hammer Waves .............................................12
`
`2.2 Method of Characteristics .......................................................................12
`
`2.2.1 Finite Difference Equations ........................................................16
`
`2.2.2 Nomenclature ..............................................................................18
`
`2.3 Boundary Conditions ..............................................................................19
`
`2.3.1 Flowrate as a Specified Function of Time at Upstream End ......20
`
`2.3.2 Series Connection .......................................................................20
`
`2.3.3 Downstream Boundary Condition ..............................................21
`
`2.3.4 Definition of Hydraulic Impedances ...........................................23
`
`2.3.5 Analogous Electrical Circuit Representation ..............................25
`
`2.4 Fracture Impedance .................................................................................30
`
`2.4.1 Fracture Dimensions from Model Parameters ............................30
`
`2.4.2 Estimate of Model Parameters ....................................................33
`
`Chapter 3: Results and Discussion .........................................................................36
`
`3.1 Hydraulic Impedance Testing .................................................................36
`
`3.2 History Matching Surface Water Hammer in Injectors ..........................38
`
`3.3 Simulated Bottomhole Water Hammer in Injectors ................................38
`
`
`
`vii
`
`IWS EXHIBIT 1060
`
`EX_1060_007
`
`
`
`3.4 Fracture Diagnostics in Injectors ............................................................51
`
`3.5 Fracture Diagnostics from Minifrac Data ...............................................51
`
`Chapter 4: Conclusion............................................................................................54
`
`Appendix ................................................................................................................56
`
`References ..............................................................................................................57
`
`
`
`
`
`viii
`
`IWS EXHIBIT 1060
`
`EX_1060_008
`
`
`
`List of Tables
`
`Table 3.1: Summary of model parameters, total surface and bottomhole pressure
`
`fluxes and attenuation. ......................................................................49
`
`Table 3.2: Summary of model parameters, equivalent fracture dimensions and near
`
`wellbore frictional pressure drop. .....................................................49
`
`Table 3.3: Comparison of fracture dimensions obtained from model with dimensions
`
`from fracture simulator for a minifrac job. .......................................52
`
`
`
`ix
`
`IWS EXHIBIT 1060
`
`EX_1060_009
`
`
`
`List of Figures
`
`Figure 1.1: Conceptual schematic of water hammer in a reservoir-pipe-valve system
`
`(From KSB Know-how series on Water Hammer) .............................4
`
`Figure 2.1: Characteristic grid in the x-t plane. .....................................................15
`
`Figure 2.2: Characteristic lines in the x-t plane. ....................................................16
`
`Figure 2.3: Nomenclature scheme for water hammer analysis. .............................19
`
`Figure 2.4: Valve opening and closing as a function of time. ...............................20
`
`Figure 2.5: Schematic of a minifrac connected to a wellbore................................26
`
`Figure 2.6: Electrical circuit representation of a minifrac. ....................................27
`
`Figure 2.7: Schematic of an injection well. ...........................................................28
`
`Figure 2.8: Electrical circuit representation of an injector. ...................................28
`
`Figure 2.9: Schematic of water hammer decline and near wellbore average pressure.
`
`...........................................................................................................30
`
`Figure 3.1: Simulated HIT for well with open fracture. ........................................37
`
`Figure 3.2: Simulated HIT for well with closed fracture. ......................................37
`
`Figure 3.3: History matching overall surface water hammer in Well A. ...............41
`
`Figure 3.4: Detailed waveform comparison of water hammer in Well A. ............41
`
`Figure 3.5: History matching overall surface water hammer in Well B. ...............42
`
`Figure 3.6: Detailed waveform comparison of water hammer in Well B. .............42
`
`Figure 3.7: History matching overall surface water hammer in Well C. ...............43
`
`Figure 3.8: Detailed waveform comparison of water hammer in Well C. .............43
`
`Figure 3.9: History matching overall surface water hammer in Well D. ...............44
`
`Figure 3.10: Detailed waveform comparison of water hammer in Well D. ..........44
`
`Figure 3.11: History matching overall surface water hammer in Well E. .............45
`
`
`
`x
`
`IWS EXHIBIT 1060
`
`EX_1060_010
`
`
`
`Figure 3.12: Detailed waveform comparison of water hammer in Well E. ...........45
`
`Figure 3.13: Misrepresentation of water hammer data due to the effect of under-
`
`sampling (after Wang et al., 2008)....................................................46
`
`Figure 3.14: Simulated bottomhole water hammer for well A ..............................46
`
`Figure 3.15: Simulated bottomhole water hammer for well B. .............................47
`
`Figure 3.16: Simulated bottomhole water hammer for well C. .............................47
`
`Figure 3.17: Simulated bottomhole water hammer for well D. .............................48
`
`Figure 3.18: Simulated bottomhole water hammer for well E. ..............................48
`
`Figure 3.19: Simulated bottomhole water hammer for a downhole shut-in in well A.
`
`...........................................................................................................50
`
`Figure 3.20: Simulated bottomhole water hammer for a downhole shut-in in well A
`
`with different formation properties. ..................................................50
`
`Figure 3.21: Comparison of modeled and measured surface water hammer pressure
`
`for a minifrac job. .............................................................................53
`
`Figure3.22: Comparison of modeled and measured bottomhole water hammer data for
`
`a minifrac job. ...................................................................................53
`
`
`
`
`
`xi
`
`IWS EXHIBIT 1060
`
`EX_1060_011
`
`
`
`Chapter 1: Introduction
`
`A change in flow causes a change in pressure, and vice-versa, which leads to
`
`transients in hydraulic systems. Water hammer is a surge or pressure wave that is created
`
`due to a sudden change in flow velocity in a confined system. It is a transient
`
`phenomenon that may be triggered by abrupt opening or closing of valves, starting or
`
`stopping of pumps, failure of mechanical devices in a flow line, etc. The name, water
`
`hammer, originates from the hammering sound that sometimes accompanies this
`
`phenomenon (Parmakian, 1963). The variation in pressure due to water hammer can be
`
`large, sometimes in the order of thousands of psi. The pressure fluctuations then
`
`propagate in the system like a wave and may cause severe damage.
`
`A conceptual schematic of water hammer in a simple system is shown in Fig. 1.1.
`
`The system consists of a frictionless horizontal pipe of constant diameter, which is fed by
`
`a reservoir at constant pressure, and is connected to a downstream valve that is suddenly
`
`closed.
`
`1. At
`
`, the pressure head is steady down the length of the pipe, as
`
`shown by the constant hydraulic grade line (shown in red), because
`
`friction was neglected, and the flow velocity is
`
`.
`
`2. As soon as the valve is shut-in, the fluid element closest to the valve
`
`comes to rest, and this rate of change of momentum causes a rise in the
`
`pressure head by
`
`. As subsequent fluid elements come to rest, the
`
`high pressure propagates upstream from the valve towards the reservoir
`
`like a pressure wave.
`
`3. At
`
`, where L is the pipe length and a is the wave speed, the high-
`
`pressure wave reaches the reservoir as all the fluid in the pipe comes to
`
`
`
`1
`
`t 0
`
`v 0
`
` H
`
`t L a
`
`IWS EXHIBIT 1060
`
`EX_1060_012
`
`
`
`rest. However, this causes a pressure discontinuity at the boundary with
`
`the constant pressure reservoir.
`
`4. In order to achieve pressure equilibrium at the reservoir, a pressure wave
`
`of magnitude
`
` is reflected back towards the valve and the direction
`
`of the flow velocity reverses towards the reservoir. This reflected wave
`
`reaches the downstream valve at
`
`. This time is called the
`
`reflection time,
`
`.
`
`5. At
`
`, the flow velocity in the entire pipe is
`
`. This causes
`
`another discontinuity at the downstream valve, where the velocity must
`
`be zero.
`
`6. The change in velocity from
`
`to zero, cause a sudden negative change
`
`in pressure of
`
`. This low-pressure wave travels upward as the fluid
`
`in the pipe again comes to rest, reaching the reservoir at
`
`.
`
`7. At
`
`, the fluid in the pipe is at rest but there is a discontinuity at
`
`the constant pressure reservoir boundary.
`
`8. As the pressure resumes the reservoir pressure, a wave of increased
`
`pressure originating from the reservoir travels back to the valve as the
`
`flow velocity in the pipe changes to
`
`.
`
`9. At
`
`, the conditions in the system are the same as 1, and the whole
`
`process starts over again.
`
`Some of the earliest studies and experiments in water hammer were done by
`
`Joukowsky (1900). The Joukowsky equation states that the rise in peizometric head
`
`(
`
`) due to the fast shut-in of a downstream valve (
`
`) is given by:
`
`
`
`2
`
` H
`
`t 2 L a
`
`T r
`
`t 2 L a
`
` v 0
`
` v 0
`
` H
`
`t 1 .5 T r
`
`t 1 .5 T r
`
`v 0
`
`t 2 T r
`
` H
`
`T c 2 L a
`
`IWS EXHIBIT 1060
`
`EX_1060_013
`
`
`
`
`
`
`
`(1.1)
`
`where a is the pressure wave-speed, V0 the initial flow velocity, g the acceleration due to
`
`gravity, L the pipe length, and Tc the valve closure time. The time period,
`
`, is the
`
`time taken by the pressure wave to propagate down the pipe length, get reflected and
`
`travel back. It is essential to consider the peak pressures due to water hammer in the
`
`design of any pipeline system, which makes water hammer a well-studied topic in civil
`
`engineering. Various researchers have simulated transient flow in pipeline systems with
`
`different methods, a discussion of which follows in the next section.
`
`Water hammer is a fast transient in the wellbore as compared to the conventional
`
`pressure transient response of the reservoir. Water hammer, in the upstream petroleum
`
`industry, has been a largely under-studied phenomenon. However, it has been a known
`
`issue following emergency shut-ins of water injectors. Due to safety concerns, the
`
`number of emergency shut downs of offshore injection wells can be high, with more than
`
`80 emergency shut downs per year in some cases (McCarty and Norman, 2006). In recent
`
`years, with the increasing number of offshore water injectors, it has been observed that
`
`injection wells that undergo repeated shut-ins show reduced injectivity, higher sand
`
`production and even failure of downhole completion (Vaziri et al., 2007). This
`
`observation has been widely attributed to the cyclic pressure waves induced by water
`
`hammer (Santarelli et al., 2000; Hayatdavoudi, 2006; McCarty and Norman, 2006; Vaziri
`
`et al., 2007; Wang et al., 2008). It is believed that in weak sands, pressure fluctuations as
`
`low as tens of psi, at the sand face, might be sufficient to cause sand failure (Santarelli et
`
`al., 2000). Modeling work by Vaziri et al. (2007) have also shown that cyclic pressure
`
`fluctuations cause more sanding than a monotonic increase in injection pressure.
`
`
`
`3
`
` H
`
`aV0
`
`g
`
`2 L a
`
`IWS EXHIBIT 1060
`
`EX_1060_014
`
`
`
`Figure 1.1: Conceptual schematic of water hammer in a reservoir-pipe-valve system
`(From KSB Know-how series on Water Hammer)
`
`
`
`4
`
`
`
`IWS EXHIBIT 1060
`
`EX_1060_015
`
`
`
`The magnitude of water hammer measured at the wellhead is often in the order of
`
`hundreds of psi, however, there is almost no bottomhole water hammer pressure data in
`
`injectors to confirm the magnitude at the sand face. It is, therefore, important to model
`
`water hammer in injectors. The first objective of this study is to model water hammer in
`
`injectors in order to estimate bottomhole water hammer pressures from measured surface
`
`data.
`
`It is also well known that sonic waves can be used to determine important
`
`information about fracture and formation properties (Mathieu, 1984; Medlin, 1991). In
`
`fact, it has been proved that fracture dimensions can be estimated from the propagation
`
`and reflection of a single pressure pulse induced at the surface of a wellbore (Holzhausen
`
`and Gooch, 1985; Paige et al., 1992). In principle therefore, it should also be possible to
`
`extract similar information from the analysis of water hammer pressure waves. Moreover,
`
`there is almost always a pressure gauge at the wellhead and water hammer pressure data
`
`can be collected without any extra effort. Hence, any information that can be derived
`
`from this data, independent of conventional testing methods, could be attractive and
`
`useful to the oil industry. The second objective of this study is to develop a model to
`
`estimate fracture connectivity and/or dimensions from water hammer pressure data.
`
`1.1 LITERATURE REVIEW
`
`1.1.1 Water hammer Modeling
`
`Classical solutions of the basic unsteady flow equations were developed by
`
`Allievi (1902, 1913) by analytical and graphical methods after neglecting the friction
`
`terms. Bergeron (1935, 1936) also developed graphical solutions that were used
`
`popularly before the advent of computers. Friction could be included by complex
`
`procedures and for practical reasons the analysis was limited to single pipelines. Streeter
`
`
`
`5
`
`IWS EXHIBIT 1060
`
`EX_1060_016
`
`
`
`and Wylie (1967) proposed and popularized the explicit method of characteristics (MOC)
`
`to solve the water hammer equations. Shimada and Okushima (1984) solved the water
`
`hammer equations by a series solution method and a Newton-Rhapson method. Chaudhry
`
`and Hussaini (1985) used MacCormak, Lambda and Gabutti Finite Difference (FD)
`
`schemes to numerically solve the water hammer equations. Izquierdo and Iglesias (2002,
`
`2004) developed a computer program using method of characteristics to simulate
`
`transients in simple and complex pipeline systems. Silva-Araya and Chaudhury (1997)
`
`solved the hyperbolic part of the equations in one-dimensional form by MOC and the
`
`parabolic part in quasi-two-dimensional using finite difference. Ghidaoui et al. (2002)
`
`proposed a two-layer and five-layer eddy viscosity model for water hammer where a
`
`dimensionless parameter (the ratio of the time scale of the radial diffusion of shear to the
`
`time scale of wave propagation) was used to estimate the accuracy of the assumption of
`
`flow axisymmetry. Zhao and Ghidaoui (2003) have solved a model for quasi-two-
`
`dimensional turbulent water hammer flow. Zhao and Ghidaoui (2004) have also
`
`developed first and second-order Godunov-type explicit finite volume (FV) schemes for
`
`water hammer problems. They have compared their schemes with MOC considering
`
`space-line interpolation for three test cases with and without friction. They found that the
`
`first-order FV schemes have the same accuracy as MOC with space-line interpolation but
`
`for a given level of accuracy, the second-order scheme requires much less memory and
`
`execution time than the first-order Godunov-type scheme. Wood (2005a, 2005b)
`
`proposed the Wave Characteristic Method (WCM) and demonstrated that though, both
`
`WCM and MOC, have the same level of accuracy, the WCM is more computationally
`
`efficient for complex pipe systems. Greyvenstein (2006) proposed an implicit FD method
`
`based on the simultaneous pressure correction approach. Afshar and Rohani (2008)
`
`proposed a water hammer simulation using an implicit MOC scheme. It is evident from a
`
`6
`
`IWS EXHIBIT 1060
`
`EX_1060_017
`
`
`
`study of the previous work done that there are various numerical models such as explicit
`
`and implicit Method of Characteristics, explicit and implicit finite difference, finite
`
`volume and finite element to solve hydraulic transient problems. Among these methods,
`
`the explicit MOC is the most popular for water hammer simulations for being simple to
`
`code, accurate and efficient.
`
`The general way of calculating friction losses in transient flows were using
`
`formulae developed for steady-state conditions, for example the use of Darcy-Weisbach
`
`equation for friction based on the mean flow velocity assumes that the shear stress at the
`
`wall is the same for steady-state and transient flow conditions. The MOC solutions were
`
`improved by incorporating unsteady or transient friction models instead of constant or
`
`steady state friction used in the early models. Zielke (1968) proposed a convolution based
`
`frequency dependepent model of unsteady friction for laminar flows that was very
`
`computationally intensive. Trikha (1975) improved the computation speed of Zielke’s
`
`model by using approximate expressions for Zielke’s weighting functions. Vardy and
`
`Brown (2004) evaluated wall shear stress in unsteady pipe flows building on the previous
`
`work by Trikha, but their solutions were faster and valid for both laminar and turbulent
`
`flows. Vardy and Hwang (1991) adopted a five-region turbulence model and a different
`
`expression in each region to compute the eddy viscosity distribution. Silva-Araya (1993)
`
`incorporated an energy dissipation factor to compute laminar and turbulent unsteady
`
`friction losses. Brunone et al. (1991) proposed a model where the total friction was the
`
`sum of a qusi-steady friction and an unsteady friction that depended on the instantaneous
`
`local and convective acceleration. Bergant et al. (2001) incorporated the two unsteady
`
`friction models by Zielke (1968) and Brunone et al. (1991) into MOC and compared the
`
`results against experiments. They found the Brunone model to be computationally
`
`effective. Saikia and Sarma (2006) proposed a numerical model using MOC and unsteady
`
`7
`
`IWS EXHIBIT 1060
`
`EX_1060_018
`
`
`
`friction calculated at every time step using Barr’s (1980) explicit friction factor
`
`correlation.
`
`Water hammer in injectors have been modeled by Moos et al. (2006) and Wang et
`
`al. (2008) as a Stoneley wave of amplitude given by Joukowsky’s formula (Eq. 1.1) that
`
`propagates down the wellbore. Moos et al. (2006) have also demonstrated that using the
`
`known physics of Stoneley wave propagation and attenuation in rocks, the formation
`
`permeability and porosity can be estimated from water hammer data.
`
`1.1.2 Fracture Impedance
`
`Khalevin (1960), Walker (1962) and Morris et al. (1964) have used acoustic
`
`waves to detect wellbore fractures. Mathieu (1984) derived analytically that Stoneley
`
`waves could be used to detect hydraulic fractures by realizing that the presence of a
`
`fracture changed the acoustic impedance of the wellbore. Mathieu derived the reflection
`
`and the transmission coefficients for waves in fractured wellbore and introduced the term
`
`“fracture impedance”. Hornaby et al. (1989) and Tang and Cheng (1989) subsequently
`
`extended Mathieu’s work to vertical and horizontal fractures. Medlin (1991) introduced
`
`tube waves (very low frequency Stoneley waves) to detect high permeability fractured
`
`zones and the connectivity of such zones with the cased hole. Holzhausen et al. (1985,
`
`1986) proposed that the altered acoustic impedance due to a fracture in the wellbore
`
`could also be demonstrated and analyzed by the characteristics of pressure oscillations at
`
`the wellhead. A single artificially induced pressure pulse at the surface would propagate
`
`down the wellbore, get reflected and be transmitted back to the surface. This Hydraulic
`
`Impedance Testing (HIT) used a lumped resistance-capacitance in series to model the
`
`fracture and estimate the fracture impedance from the reflected pulse by trial and error.
`
`The amplitude of the reflected pulse, which is determined by the impedance contrast
`
`
`
`8
`
`IWS EXHIBIT 1060
`
`EX_1060_019
`
`
`
`between the wellbore and the fracture, could be used to compute the width and height of
`
`the fracture. Holzhausen’s model was experimentally validated by Paige et al. (1992) and
`
`some field scale tests were also carried out (Paige et al., 1993; Holzhausen and Egan,
`
`1986). Paige et al. (1992) showed that the pressure wave would reach the tip of the
`
`fracture and proposed that the length of the fracture could be estimated, by measuring the
`
`time lapse between the reflections of the wave from the entrance to the fracture and the
`
`tip of the fracture. Ashour (1994) generalized Holzhausen’s HIT method for vertical and
`
`horizontal hydraulic fractures and showed that sending a wave that is close to the
`
`resonance frequency of the fracture can make a more accurate assessment of fracture
`
`dimensions. Holzhausen’s model assumed no energy losses in the wellbore, which meant
`
`that the attenuation of the pressure wave due to friction in the wellbore was not taken into
`
`consideration. Patzek and De (2000) overcame this issue by proposing a lossy
`
`transmission line model to describe the wellbore and fracture geometry and capture the
`
`wellbore and fracture dynamics. In their model, flow through both the wellbore and the
`
`fracture was treated analogous to the flow of electricity through transmission lines and
`
`resistance, capacitance and inductance were distributed over the length of the line.
`
`However, it has been the general opinion that it is difficult to collect the required
`
`information from the measured pressure signal and HIT has not been used very popularly
`
`in the industry.
`
`
`
`9
`
`IWS EXHIBIT 1060
`
`EX_1060_020
`
`
`
`Chapter 2: Model Formulation
`
`2.1 WATER HAMMER MODELING EQUATIONS
`
`The basic differential equations for transient flow in closed conduits are the one-
`
`dimensional conservation equations of mass (continuity equation) and momentum
`
`(equation of motion). The generalized forms of these equations were derived by using the
`
`Reynolds transport theorem and then simplified using assumptions that are valid for
`
`water hammer analysis (Chaudhury, 1987). Wylie and Streeter (1993) have also provided
`
`a detailed derivation and discussion of these governing equations. The following sections
`
`present a brief description of these equations from their work.
`
`2.1.1 Continuity Equation
`
`The general form of the continuity equation is
`
`
`
`
`
`(2.1)
`
`where, A = area of cross-section of conduit, ρ = density of the fluid, V = mean flow
`
`velocity, t = time, x = coordinate axis along the axis of the conduit. The first term in Eq.
`
`(2.1) accounts for the compressibility of the fluid and the second term represents the rate
`
`of deformation of the conduit wall.
`
`
`Assuming an elastic conduit filled with a slightly compressible fluid, Eq. (2.1)
`
`simplifies to
`
`
`
`
`
`(2.2)
`
`where, p = pressure intensity, V = mean flow velocity, a = wave speed or the velocity of
`
`the water hammer waves. For low-Mach-number unsteady flows, the transport term
`
`
`
`10
`
` 0
`
`
`
`V
`
`x
`
`dA
`
`dt
`
`1 A
`
`
`
`d
`
`dt
`
`1
`
`p
`
`t
`
` V
`
`p
`
`x
`
` a 2 V
`x
`
` 0
`
`IWS EXHIBIT 1060
`
`EX_1060_021
`
`
`
` is small compared to the other terms and may be dropped to yield the simplified
`
`continuity equation
`
`
`
`2.1.2 Equation of Motion
`
`
`
`(2.3)
`
`The general form of the momentum equation is
`
`
`
`
`
`(2.4)
`
`where, f = Darcy-Weisbach friction factor, θ = angle of inclination of the pipe and, D =
`
`diameter of the pipe.
`
`Once again, the convective transport term,
`
`is neglected for low-Mach-
`
`number unsteady flows, reducing Eq. (2.4) to
`
`
`
`
`
`(2.5)
`
`It is often convenient to analyze pipeline flows by defining pressure, p, in terms of
`
`the piezometric head, H and use the discharge, Q, instead of the flow velocity, V.
`
`
`
`
`
`(2.6, 2.7)
`
`where, p = pressure, g = acceleration due to gravity, ρ = density of the fluid, z = elevation
`
`of the pipe above a specified datum, V = mean flow velocity, and, A = area of cross-
`
`section of the pipe.
`
`Eqs. (2.3) and (2.5) expressed in terms of H and Q, become:
`
`
`
`11
`
`V p x
`
`p
`
`t
`
` a 2 V
`x
`
` 0
`
`fV V
`
`2 D
`
` 0
`
` g sin
`
`p
`
`x
`
`1
`
`
`
`V
`
`x
`
` V
`
`V
`
`t
`
`V V x
`
`fV V
`
`2 D
`
` 0
`
` g sin
`
`p
`
`x
`
`1
`
`
`
`V
`
`t
`
` z
`
`p
`
`g
`
`H
`
`Q V A
`
`IWS EXHIBIT 1060
`
`EX_1060_022
`
`
`
`
`
`
`
`(2.8, 2.9)
`
`2.1.3 Velocity of Water Hammer Waves
`
`The following general expression for the wave propagation velocity a, was
`
`presented by Halliwell (1963)
`
`
`
`
`
`(2.10)
`
`where, ψ = nondimensional parameter that depends on the elastic properties of the
`
`conduit, E = Young’s modulus of elasticity of the conduit walls, K = bulk modulus, and ρ
`
`= density of the fluid, respectively.
`
`Expressions of ψ under various conditions (rigid conduit, thick-walled elastic
`
`conduits, thin-walled elastic conduits, tunnels through solid rock, reinforced concrete
`
`pipes, etc.) are available in the literature (Chaudhry, 1987; Wylie and Streeter, 1993). For
`
`our analysis, we use the expression of ψ, valid for thin-walled elastic conduits anchored
`
`against longitudinal movement throughout its length, given as
`
`
`
`
`
`(2.11)
`
`where, D = conduit diameter, e = wall thickness, ν = Poisson’s ratio of pipe material.
`
`2.2 METHOD OF CHARACTERISTICS
`
`The water hammer modeling equations are a pair of quasi-linear, hyperbolic,
`
`partial differential equations and a closed-form solution of these equations is not
`
`
`
`12
`
`H
`
`t
`
`
`
`a 2
`
`Q
`
`g A
`
`x
`
` 0
`
`Q
`
`t
`
` g A
`
`H
`
`x
`
`
`
`fQ Q
`
`2 D A
`
` 0
`
`a
`
`K
`
`
`
` 1 K E
`
`
`
`
`
`
`1 2
`
`
`
`D e
`
`
`
`IWS EXHIBIT 1060
`
`EX_1060_023
`
`
`
`available. However, there are several methods to numerically integrate these equations,
`
`such as, method of characteristics, explicit and implicit finite-difference methods, finite-
`
`element methods, etc. Amongst these, the method of characteristics has been the most
`
`popular due to its several advantages over other methods, particularly in water hammer
`
`type problems. These advantages include an explicit form of solution such that different
`
`elements can be solved independently and complex pipe networks can be handled with
`
`ease, an established stability criterion, an easy to program and computationally efficient
`
`procedure and most importantly, accurate solutions. The main disadvantage of this
`
`method is the requirement to adhere to the time step-distance interval relationship.
`
`The momentum and continuity equations, in terms of two dependent variables,
`
`discharge and piezometric head, and two independent variables, distance along the pipe
`
`and time, are transformed into four ordinary differential equations by the method of
`
`characteristics. For further discussion let us rewrite the momentum and continuity
`
`equations (Eqs. 2.8 and 2.9) as
`
`
`
`
`
`
`
`(2.12, 2.13)
`
`A linear combination of these equations using an unknown multiplier λ yields
`
`
`
`(2.14)
`
`Please note that, using any two real, distinct values of λ, Eq. (2.14) will again
`
`yield two equations that are equivalent to Eqs. (2.12) and (2.13). Also, if
`
`
`
`and
`
`, then the total derivative can be written as
`
`
`
`13
`
`L1
`
`Q
`
`t
`
` g A
`
`H
`
`x
`
`
`
`fQ Q
`
`2 D A
`
` 0
`
`L 2 a 2 Q
`
`x
`
` g A
`
`H
`
`t
`
` 0
`
` 0
`
`fQ Q
`
`2 D A
`
`
`
`
`
`H
`
`t
`
`1
`
`
`
`H
`
`x
`
`
`
` g A
`
`
`
` a 2 Q
`x
`
`Q
`
`t
`
`
`
`L L1 L 2
`
`H H ( x, t )
`
`Q Q ( x, t )
`
`IWS EXHIBIT 1060
`
`EX_1060_024
`
`
`
`
`
`
`
`
`
`
`
`(2.15, 2.16)
`
`It can be seen by from Eqs. (2.14), (2.15) and (2.16), that if λ is defined as
`
`
`
`(2.17)
`
`Then, by substituting the two particular values of λ, Eq. (2.14) can be written as two pairs
`
`of equations and identified as
`
`and
`
` equations.
`
`
`
`
`
`(2.18, 2.19)
`
`Thus, by imposing a relationship between the two independent variables, the
`
`original partial differential equations (Eqs. 2.8 and 2.9) were converted to two total
`
`differential equations. These ordinary differential equations, however, are not valid
`
`everywhere in the x-t plane like the Eqs. (2.8) and (2.9) were. Instead, Eq. (2.18) and Eq.
`
`(2.19) is only valid alo