You are on page 1of 27

PE 7013

Advanced Reservoir Engineering

(c) Rami M. Younis, 2023

August 30, 2023


Chapter 1

Governing Equations

Reservoir Engineers are concerned with asset valuation, development planning, and designing
operations to improve production performance. Ultimately, all of these activities require the ability
to model the reservoir and petroleum fluid movement within it given certain operating conditions.
There is basically only one governing principle that reservoir engineers need to apply; continuity.
Continuity describes the transport of a conserved quantity such as mass, energy, momentum, and
electric charge. The two conserved quantities that are of current interest to petroleum reservoir
engineers are mass and momentum. The conservation of mass governs pressure and chemical species
distributions under a pressure driving force. The conservation of momentum governs the dynamic
of motion of the rock as it deforms due to the action of geomechanical forces and pore pressure.
Combined with constitutive relations, the general continuity equation provides a forward model
for the spatiotemporal dynamics of state variables. Reservoir engineers will also perform volume
averages of continuity in order to study the temporal evolution of volume averaged state variables
(Material Balance Equations). This chapter introduces general continuity and its volume averaged
form. We then specialize to the conservation of mass and momentum. Finally, through problems
at the end of the chapter, the art and challenge of formulating well-posed governing equations for
physical systems is explored.

1.1 Continuity
The continuity equation describes the balance of a conserved quantity, [M (x, t)] = Θ over
space and time. There are two forms of the same continuity statement. The first form arises
when considering a control volume, Ω ⊂ RN , where N = 1, 2, or 3 is the dimension of the spatial
domain. The control volume is assumed to have a boundary denoted ∂Ω ⊂ RN −1 (which is not
necessarily connected if for example there are wholes inside the domain). At any point on the
boundary y ∈ ∂Ω, the outward pointing unit normal vector is assumed to be well-defined and is
denoted n̂ (y). In this form, the continuity statement is an overall balance over the entire domain,
and the whole picture is summed up over space by integral forms of the statement. The second
form is a differential form that governs continuity at every point in space. Clearly, since the two
statements are equivalent, one may derive one from the other. We will start by deriving or stating
the integral form and producing the differential form from it.

The point of departure is to define a spatial distribution of the volumetric density of the quan-
tity to be conserved as the scalar field, m (x, t). This field is assumed to be well-defined throughout

1
Figure 1.1: A control volume Ω with boundary ∂Ω and outward point unit normal n̂ (x) is immersed
in a flow field q (x).

the control volume of interest, Ω. If the units of the conserved quantity are Θ, then the units of
this bulk volume density field are [m] = Θ.L−N . We will also assume that the conserved quantity
is transported in space and time according to the flux vector field q (x, t). Flux is defined as the
rate of transport of the conserved quantity per unit area, and the units are [q] = Θ.L1−N .T −1 . In
addition to the flux field, there may be some sources or sinks that inject or extract the quantity
at given points. In general, this source field will be denoted as s (x, t) with units of ΘL−N T −1 .
Figure 1.1 depicts this whole situation in general.

Before deriving the continuity equation, one more consideration must be accounted for. It is
possible that the domain representing the bulk volume is moving or deforming in time. Given the
initial spatial configuration of the domain, Ω0 , the time-dependent domain is defined as,
Ω (t) = {x (t) : x (0) ∈ Ω0 } ,
and the velocity of every point x (t) ∈ Ω (t) in the domain is specified as,
dx
c= (t) .
dt
Note that if the domain were stationary, then Ω (t) = Ω0 and c = 0. It is also important to note
that c is the bulk velocity of the whole domain, and that the flux field, q, is relative to this velocity
with respect to a statoinary point of view.

The conservation principle stated in simple terms is that the rate of change of the total con-
served quantity in the control volume must equal the net rate of influx of the conserved quantity

2
through the boundary plus the rate of influx due to sources and sinks. We will write out these
three terms individually first and then combine them into the continuity equation.

Mathematically, the rate of change of the total conserved quantity is simply,


Z
dM ∂
= m (x, t)dx (1.1.1)
dt ∂t Ω(t)
Z Z
∂m
= (x, t)dx + m (x, t) (c · n̂)dx (1.1.2)
Ω(t) ∂t ∂Ω(t)
Z
∂m
= (x, t) + ∇ · (m (x, t) c)dx (1.1.3)
Ω(t) ∂t

Notice the application of the identity developed in Appendix A (A.4.1) in order to differentiate
a volume integral in time, followed by an application of the Gauss-Green Theorem that was also
presented in Appendix A to convert a surface integral into the volume integral of the divergence.

Next we consider the rate of total influx of the quantity across the boundary, Q. At any point
on the boundary x ∈ ∂Ω, the flux into the volume is given by the inner product, −q (x, t) · n̂ (x).
Subsequently, the total influx on mass is given by integrating over the boundary:
Z
Q= −q (x, t) · n̂ (x)dx (1.1.4)
∂Ω(t)
Z
= −∇ · q (x, t)dx (1.1.5)
Ω(t)

Finally, the rate of total injection due to the sources and sinks, S is simply,
Z
S= s (x, t)dx.
Ω(t)

The continuity equation is now easily obtained by equating the first term with the sum of the
second two;
dM
(t) = Q (t) + S (t)
dt
This can be accomplished using any mix of the volume or surface integral options for dM dt
and
Q. They all lead to the same statement of continuity and depending on the application at hand,
there may be advantages to using one form over the other. For example, using surface integrals,
we can write the continuity equation as,
Z Z Z
∂m
(x, t)dx = s (x, t)dx − (m (x, t) c + q (x, t)) · n̂dx (1.1.6)
Ω(t) ∂t Ω(t) ∂Ω(t)

whereby, given the information on the boundary, the boundary integral can be evaluated as data,
allowing to solve an integral equation for the rate of change of the total quantity. Alternatively,
choosing the volume integral forms with the expanded time derivative, we have the form,
Z
∂m
(x, t) + ∇ · (m (x, t) c) + ∇ · q (x, t) + s (x, t)dx = 0. (1.1.7)
Ω(t) ∂t

3
In the above form, the volume integral acts to sum up the pointwise continuity statement. In
fact since for any physical system, we can choose an inifinite number of domains for which the
statement must hold, then it must also hold pointwise. Another way is to choose Ω arbitrarily
small, and the statement would still need to hold. Therefore, the following differential version
holds at avery point in space regardless of its position within or outside of any control volume:
∂m
(x, t) + ∇ · (m (x, t) c) + ∇ · q (x, t) + s (x, t) = 0. (1.1.8)
∂t
The differential form is a profound statement that holds pointwise in any continuum. It is most
often applied to help us to determine the spatiotemporal evolution of the density variable. When
combined with constitutive relations for the density and flux in terms of the thermodynamically
independent variables, then they may be determined as well.

1.1.1 Auxilliary conditions


Recall from Section ?? the idea of a finite number of independent state variables from which all
other state variables can be derived. In a given system, the density distribution and flux fields may
be functions of the indpendent state variables. Moreover additional constitutive relations such as
Darcy’s law may help us to specify the flux in terms of the independent variables. Therefore, if
the continuity equations could be inverted in order to obtain the spatiotemporal evolution of the
independent state variable, we may learn about all other aspects of the system. This will often be
the case in reservoir engineering where essentially continuity of mass, energy, and momentum are
all that are required to pin down any system! One piece to the puzzle remains however, and that
is, what data or measurements and observations must be specified in order for continuity to be
well-posed? Well-posedness means that the equations are invertible and that the solution is unique.

Equations 1.1.6, 1.1.7, and 1.1.8 all have a single time derivative. Therefore additional data
must be provided to pin down the temporal origin of the solution manifold. This is typically in
the form of an initial condition. The initial condition will typically provide data or measurements
of the independent state variable field at some point in time. This may be a direct measurement
or an indirect one that can itself be inverted.

Moreover, both integral and differential forms require measurements of the flux and deformation
velocity at the boundary of the domain. In Equation 1.1.6 these are required to calculate the
boundary integral. In the differential form, these complement the divergence operator.

1.1.2 Coordinate systems and symmetry


Notice that the three conservation equation forms do not explicitly say anything about the
coordinate system that is being used or the number of dimensions involved. It is general to any
system you are working in provided that you can specify definitions for the divergence and the
gradient operators. In Cartesian coordinates, we have,
∂p ∂p ∂p
∇p = ix + iy + iz ,
∂x ∂y ∂z
and,
∂ux ∂uy ∂uz
∇.u = + + ,
∂x ∂y ∂z

4
for the gradient and divergence operators respectively. Appendix A (A.2.2) explains how to write
the divergence and gradient operators in other coordinate systems.

Often reservoir engineers will talk about symmetry in a problem. Symmetry refers to cases
where it is known ahead of time that the independent state variables only vary in one or two
spatial coordinates rather than three. When this happens, while the physical dimensionality of
the problem is always three, the dimensionality of the mathematical problem reduces to onw or
two. Symmetry is specific to the coordinate system being used. For example a radially symmetric
system in cylindrical coordinates is one dimensional mathematically whereas it is two dimensional
in the Cartesian frame.

1.1.3 Macroscopic governing equations


The microscopic view assumes that we are interested in the spatiotemporal distribution of state
variables. Often in engineering applications, we want to study the time evolution of the average
properties in a reservoir. The macroscopic equations aim to do this by introducing volume averaged
quantities. Referring to Figure 1.1, the total bulk volume of the reservoir is,
Z
Vb = 1dx.

We introduce the concept of the volume averaged continuum quantities. For the density field, this
is defined as R
m (x, t)dx
m (t) = Ω
Vb
Subsequently, we might consider the volume averaged from of the continuity equation,
Z
∂ 1
m (t) = − (q (x, t)) · n̂dx (1.1.9)
∂t Vb ∂Ω(t)
which may provide us with the discrete balance form,

1 t
Z Z
m (t) − m (0) = − (q (x, t)) · n̂dxdt (1.1.10)
Vb 0 ∂Ω(t)

1.1.4 Formulating well-posed governing equations


Given practical problems of interest for which there are no cookie-cutter results or well-studied
rules of thumb, the advanced engineer is able to quickly formulate a mathematical model that
governs the system. An advanced and talented engineer will do so providing a well-posed system
that is easily solved. It takes a great deal of practice to become adept and often there is a need
for creativity and common sense.

While there is no universal procedure one can (or ought to) follow, the following is a subjective
modeling framework that can be used.
1. Sketch a physical or mental picture of the system and imagine what is happening over time.
Use your intuition. This may not always be possible, and in fact, for worthy problems where
we have very little experimental experience this may be misleadingly wrong. In such cases,
the element of creativty becomes important.

5
2. Identify the independent state variables that are necessary for you to completely determine
the evolution of the system. Are you interested in averages of these quantities over a domain,
or the pointwise spatial distribution? How many independent state variables do you have?
3. Identify control volumes that are undergoing various processes. There may be more than
one, and these systems may be coupled through a specific transfer function.
4. For each control system, identify the continuity equations that must hold. For each one,
determine the denisty and flux functions and use constitutive relations to express them in
terms of the independent state variables.
5. for each continuity equation, think about what observations or data is available before hand.
Determine sufficient auxilliray (initial and boundary) conditions to close the system.

1.2 Conservation of mass


In this section the continuity equation is applied to govern the conservation of mass. Mass
(chemical species or psuedo-components) must be conserved. A the mass of a phase is not phys-
ically conserved unless phase changes are intrinisically modeled as in the Black oil model with
solution gas or volatile oil. Each chemical or conserved phase will require its own continuity equa-
tion. In this section, we will consider non-deforming rock that is stationary. Moreover, we will
assume isothermal processes meaning that the temperature field is constant in time. If we did not
assume these two things, we will see that we will also need to write conservation of momentum
and energy equations as well in order to have a well-posed system.

1.2.1 Single phase flow


When considering the flow of a single phase in rock, the single phase can essentially be thought
of as one chemical component. According to the Gibb’s phase rule, we can consider pressure p (x, t)
and temperature T (x, t) as the set of independent variables. Since the problem is isothermal, we
will neglect temperature as an explicit variable and drop it from the notation for brevity. We will
also neglect source-sink terms although they are easily incorporated.
Subsequently, the mass density of fluid per unit bulk volume is simply, m (p) = φρ where φ (p)
and ρ (p) are the pressure dependent porosity and fluid density. The flux is then simply the fluid
density time the Darcy velocity:
q (p) = ρu
ρ
= − k (x) ∇ (p − ρd (x))
µ
where d (x) is depth along gravity at any position x. Then the first form of the conservation
equation corresponding to Equation 1.1.6 reads,
Z Z
∂ ρ
φρdx = [K (x) .∇ (p − ρd (x))] · n̂ (x)dx,
Ω ∂t ∂Ω µ
and the PDE form (corresponding to Equation 1.1.8) reads,
 
∂ ρ
φρ = ∇ · K (x) .∇ (p − ρd (x)) .
∂t µ

6
1.2.2 The black oil model
In the black oil model, the introduction of the formation volume factors and solution gas ratio
and volatile oil ratio allow us to conserve surface volumes of fluid since the surface density is
constant. In this general model, assuming isothermal conditions, there are three independent
variables: po (x, t), Sw (x, t), and So (x, t). Given these three variables, we can caluclate all other
variables using thermodynamic property relations and the following constitutive constraints:

Sg = 1 − So − Sw ,
pw = po − P cow (Sw ) , and,
pg = po + P cog (So ) .

The Darcy velocity for each phase ν ∈ {w, o, g} is written as,

Krν
uν = − K (x) .∇ (pν − ρν d (x)) (1.2.1)
µν
Subsequently, the conservation of stock-tank oil, surface gas, and stock-tank water are simply;
   
∂ So Sg Rv 1 Rv
φ + +∇· uo + ug = 0
∂t Bo Bg Bo Bg
   
∂ So Rs Sg Rs 1
φ + +∇· uo + ug = 0
∂t Bo Bg Bo Bg
   
∂ Sw 1
φ +∇· uw = 0
∂t Bw Bw

By integrating over the volume of the reservoir and over a time interval, these equations can
be expressed in terms of the average properties as,
 
So Sg Rv
Vb φ + − N = 0 − Np
Bo Bg
 
So Rs Sg
Vb φ + − G = GI − Gp
Bo Bg
 
Sw
Vb φ − W = WI − WP
Bw
These are the so-called Material Balance Equations that are the subject of an undergraduate
introductory course in reservoir engineering.

1.3 Conservation of energy


The conservation of energy is particularly important for non-isothermal processes such as ther-
mal EOR or downhole heaters. We will develop the appropriate terms for the density and flux field
as well as the source terms. We will once again assume a non-defomrable rigid rock although it is

7
intructive to think about how things will change in the case that this assumption is not made. We
will also assume that the rock is saturated with a multi-phase system where there are N p phases
in the index set ν ∈ Ip.

The energy density per unit bulk volume depends on the amount of internal energy in the
system. Since the rock and the fluid can carry this internal energy, we have,
X
m=φ [ρν Sν eν ] + (1 − φ) ρr er
ν∈Ip

where eν is the specific internal energy of the phase at hand.

Next is the flux field. Energy can be transported by several phenomena. First, energy is carried
by mobile fluid in the form of enthalpy. It may also by transported through conduction. Radiation
is an unlikely culprit and is typically neglected. Using Fourrier’s law of conduction, and introducing
the specific phase enthalpy, the flux filed is,
X
q= [ρν hν uν ] − KT (x) .∇T
ν∈Ip

Finally sources and sinks of energy include exothermic or endothermic chemical reactions and
downhole heaters modeled as point sources (a wellbore with effectively zero radius. Wellbores with
finite radius are modeled with internal boundaries and incorporated as boundary conditions). We
will denote this total source term as sT .

The differential form of the continuity equation becomes,


" #
∂ X X
φ [ρν Sν eν ] + (1 − φ) ρr er + ∇ · [ρν hν uν ] − ∇. [KT (x) .∇T ] − sT = 0. (1.3.1)
∂t ν∈Ip ν∈Ip

1.4 Conservation of momentum


The conservation of momentum in porous media mainly relates to the coupling between flow
and geomechanics. Deforming reservoirs due to declining average pressure and large overburden
forces can lead to compaction and subsidence. In unconventional reservoirs, unpropped fractures
may experience dramatic changes in aperature and propped fractures can even experience a crush-
ing effect on the proppant. The conservation of momentum relates deformation to pore pressure,
body forces, and stress. More generally however, the conservation of momentum may be consid-
ered for rigid body dynamics or free fluid flow in cavities for example. When there is no significant
deformation and the interplay between stress and pore pressure is negligible, then the conservation
of momentum may be neglected similar to the continuity of energy for isothermal problems.

1.4.1 General momentum balance


First we consider a general control volume Ω ⊂ RN that is not a saturated porous media.
Rather it is a collection of mass that is assumed to have a mass density ρ and is subjected to a

8
stress tensor field σ ∈ RN × RN where σi,j has units M L2−N T −2 (which are the same units as those
for flux in this case!). Finally, a body force (namely gravity) acts as a source density funtion on
the body. Momentum is the conserved quantity, Θ = M LT −1 , and its density field is,
dx
m=ρ .
dt
Notice that the density field is an N -dimensional vector! This is because there will be N momentum
balance equations; one each orthogonal direction. Subsequently, the flux field must now be a tensor
field. It is in fact equal to the stress field that the mass is emmersed in,

Q = σ.

Finally the gravitational body force distribution is the vector field,

s = ρg

. We are now ready to substitute these terms into the continuity equation. First, in the surface
integral form, we have,
Z Z Z
∂ dx
ρ dx = ρgdx − σ · n̂dx
∂t Ω(t) dt Ω(t) ∂Ω(t)

If we consider a rigid body that is not losing mass, then we retrieve the famous Newton’s law,

d2 x X
M 2 = Mg − f,
dt surface

where in this context M stands for the total mass of the rigid body. How does this expression
change if there was mass loss to the environment? How about for deformable solids?

9
1.5 Worked problems
Problem 1
THE SETUP:
Consider the new patent pending device in Figure 1.2. The device consists of a spring-
loaded piston-cylinder assembly. The cylinder which is of length 2L and cross-sectional
area A = 1, is half full of incompressible water and the other half has a tightly fitted
compressible core that is saturated with incompressible water. The water density, ρ,
water viscosity, µ, and the core isotropic homogeneous permeability, k, are all con-
stant. Defining a constant isothermal compressibility, cr , the porosity of the core is
φ (x, t) = cr p (x, t) + 0.3.

THE EXPERIMENT:
Initially, the spring is compressed and the piston which is of mass, M = 1, is pulled to
its starting point at x = 0. At this point, the force in the spring along the x-direction
is σ (1 − x). The initial pressure in the system is at equilibrium and is p0 . At time,
t = 0, the triggering device is activated. The piston advances to the right into the void
space gradually until it slows down and stops just ahead of the core.

Figure 1.2: A schematic of a ”Linear Compressometer”.

(a) Suppose that the piston reaches an equilibrium position x∗ = 0.75 at t∗ = 1. Derive
an expression for the pressure in the entire cylinder at equilibrium, p∗ .

At equilibrium there is no flow anywhere. This means that the pressure gradient is zero ev-
erywhere. In turn, this means that the pressure is constant throughout the entire cylinder. At

10
equilibrium, a force balance on the piston should read,
X
F = σ (1 − x∗ ) − p∗ A = 0

Plugging in the numbers for all of the known quantities, we get,

p∗ = 0.25σ

(b) Suppose that a pressure gauge is fitted into the void part of the cylinder right at
the core’s left face. During the experiment, 0 ≤ t ≤ t∗ , the pressure is observed to
change linearly in time from p (x = 1, 0) = p0 to p (x = 1, t∗ ) = p∗ . Derive an expression
for the position of the piston as a function of time, x (t). The expression must involve
only known quantities. Use σ = 1 and p0 = 1.

The pressure in the water-filled part of the cylinder is constant; instantaneous equilibrium. This
pressure is precisely the pressure on the left face of the core; pL (t) = p (1, t). We are told that the
pressure varies linearly from 1 to p∗ = 0.25σ. The equation for a straight line is simply,
p∗ − p0
pL (t) = t + p0 = 1 − 0.75t
t∗ − 0
Now that
P we have expressions for the forces on either side of the piston, we can write its equation
of motion F = M A as follows;
d2 x
σ (1 − x (t)) − pL (t) A = 0.75t − x (t) =
dt2
The equation of motion is a second order Ordinary Differential Equation (ODE). Using the
data provided in the question, we have boundary data available to us. In particular, at the start
we have,
x (0) = 0,
and at equilibrium we have,
x (1) = 0.75.
In summary, we need to solve the ODE,



d2 x

 dt2 + x − 0.75t = 0







 x (0) = 0





x (1) = 0.75

From basic fundamental ODEs, the solution to this problem is,

x (t) = C1 cos (t) + C2 sin (t) + 0.75t.

We apply the Boundary Conditions to obtain that C1 = C2 = 0, giving us

x (t) = 0.75t.

11
(c) Derive an expression for the mass flow rate of water into the core, Qw , in terms of
known quantities only

Since the water is incompressible, the mass flow rate into the core is simply equal to the rate
dm
of change of mass in the water part of the cylinder; dtcyl = −Qw . So,
dx
Qw = −ρA = −0.75ρ.
dt
(d) Write the governing equation and the initial and boundary conditions for the mass
balance in the core. Is your formulation to this problem unique? Is the solution to
this problem unique? Explain your answers.

We can write at least two formulations; one in terms of a rate boundary condition and the other
in terms of a pressure condition. That is, in terms of rate,



k ∂2p
 ∂p
 ∂t − cR µ ∂x2 = 0








 ∂p (1, t) = 0.75 µ

∂x k



 ∂p


 ∂x
(2, t) = 0






p (x, 0) = 1

On the other hand, in terms of pressure, we have,





∂2p
 ∂p
− cRkµ ∂x2 = 0




 ∂t





p (1, t) = 1 − 0.75t



 ∂p


 ∂x
(2, t) = 0






p (x, 0) = 1

Each of these two systems is well-posed; has a unique solution. Both systems are compatible
since the rate condition was derived from the pressure condition via parts (b) and (c). So the
solutions should also be compatible.

Problem 2.
Recently, a new device was invented to measure the equilibrium pressure in a torus
shaped balloon using only measurements of the balloon radius R. The device cross-
section is depicted in Figure 1.3.

12
Figure 1.3: Cross-section of annular device. The balloon on the outside is filled with air. The disc
is a porous medium.

The device consists of a porous media disc of thickness H, porosity φ, outer radius ro ,
and permeability k. An air injection inlet is drilled into the center of the disc with
radius rw . Air of constant viscosity µ and density ρ (incompressible) is injected at a
constant pressure pw as the outer annular balloon expands. The inventor measures
the radius of the balloon as a function of time and it turns out to be R (t).

a) Derive a pressure-derivative outer-boundary condition for the flow in the disc.

This derivation is accomplished by performing a trivial mass balance on the balloon.

First, the rate of accumulation of air in the balloon is the time derivative of the total mass in
the balloon. The total mass of air in the balloon at any given time is given by ρV , where V is the
inflated volume. Note that the volume of the balloon for a given radius is simply the cross-sectional
area times the perimeter of the torus centerline,

V = πR2 .2π (ro + R) .

Next, the flow rate of mass into the balloon is simply the radially symmetric flow rate of air
coming in from the disc;

q (r = r0 ) = ρAcross ur (r = ro )
 
k
= ρ (2πro H) − ∇p (r = ro )
µ
 
ρk ∂p
= −2πro H (r = ro )
µ ∂r
We can simply equate the rate of change of accumulation to the net influx and integrate both
sides of the equation as follows,
Z t Z t
∂V
dν = q (r = r0 ) dν.
0 ∂ν 0

The integral on the left is trivial. On the other hand, since the flow is incompressible, the radially
symmetric flow rate of air in the disc is constant. This means that the integral on the right

13
simplifies. Performing the integral, we obtain;
0 = ρV (t) − ρV (0) − q (t − 0)
 
2 2 2 2 k ∂p
= 2ρπ R (ro + R) − 2ρπ R (0) (ro + R (0)) + 2ρπro H
µ ∂r r=ro
Finally, this simplifies to the pressure derivative boundary condition,
µ R2 (ro + R) − R02 (ro + R0 )
 
∂p
=π =γ
∂r r=ro Hk ro
b) Solve for an expression for the pressure distribution in the porous media disc as a
function of radius, r. Is this solution time-dependent?

The problem to solve is that of incompressible radial flow with a Dirichlet inner boundary and
a Neumann outer boundary condition:



1 ∂ ∂p
 
r =0



 r ∂r ∂r



 p (r = rw ) = pw





 ∂p 

∂r r=ro

Integrating the differential equation, we obtain the general solution,
p (r) = C1 ln r + C2 .
To determine the integration constants, we apply the boundary conditions. Let us start with the
outer condition:  
∂p
γ= r = C1
∂r r=ro
Next, the inner condition gives;
pw = γ ln rw + C2
This implies that C2 = pw − γ ln rw , or that p (r) = γ ln rrw + pw .

c) Using a force balance and the elastic properties of the balloon material, the inven-
tor comes up with a new relation R (pballoon ) = σpballoon . Derive an expression for the
pressure in the balloon pballoon as a function of σ and the terms in your answer to part
(a).

Substituting the new expression for R into p (r = ro ) and assuming for simplicity that R0 = 0,
we get,
πµσ 2 p2b (ro + σpb ) ro
pb = ln + pw
Hk ro rw
Rearranging, we get the cubic equation,
1
p3balloon + ro p2balloon − (pballoon − pw ) = 0,
α
πµσ 2
where α = Hkro
ln rrwo .

14
Problem 3.
Consider the sequence of events depicted in Figure 1.4. Starting from
the top image, a spring loaded piston-cylinder system of cross sectional
area A = 1 is at rest, i.e. the force in the spring is zero. The spring is
then compressed by a length L, and an air saturated porous medium
is inserted into the cylinder. The air is compressible, as is the porous
medium. At time t = 0, the piston is set free. It starts to move to the
left, compressing the porous medium while allowing a flow rate Q (t)
to take place. Assume that the pressure is uniform throughout the
cylinder.

Figure 1.4: Spring-loaded porous media syringe.

(a) Derive an equation for the conservation of mass in the system.

d x(t)
Z
0= φ (t) ρ (t)dν + Q (t)
dt 0
 
d
= φ (t) ρ (t) [x (t) − 0] + Q (t)
dt
= φ0 ρx + φρ0 x + φρx0 + Q (t)
= (c + c ) e(cf +cr )P (t) xP 0 + e(cf +cr )P (t) x0 + Q (t)
f r

(b) Derive an equation the force balance on the piston assuming its mass
is M = 1.

0 = M.x00 − P.A + K.x


= x00 + Kx − P (t)

15
(c) Write the coupled system of differential equations in the form Au00 +
Bu0 + Cu + D = 0, where u = (x, p)T and the coefficients are matrices.

     
 0 0  00
 u + φ (t) ρ (t)  (cf + cr ) x 1  u0 +  0 0  u = 0
   

     
0 1 0 0 −1 K

(d) Provide appropriate initial conditions.

 
 P0 
u (0) = 



L

Problem 4.
In this problem, we will consider the flow of an ideal (real) gas
through a homogeneous system. Recall that the ideal gas law is,

pV = ZRT,

where the real gas compressibility factor Z depends on pressure, R is


the universal gas constant, and T is temperature. Moreover, recall that
from basic thermodynamics, the isothermal fluid compressibility obeys
the relation,  
1 ∂V
cf = − .
V ∂p T
We will derive the pressure equation in radial coordinates for this prob-
lem.

(a) Using the ideal gas law, write an expression for the fluid density, ρ,
given a mass of gas, M .

Straight from basic definitions we have,


M Mp
ρ= = .
V ZRT

(b) Express the isothermal compressibility, cf , as a function of Z (p) and p.

16
Using the ideal gas law in the isothermal compressibility equation, we get,
 
p 0 ZRT 1
cf = − Z RT −
ZRT P P
0
1 Z
= −
p Z

(c) Using the results in (a) and (b), write out and simplify the isothermal
pressure equation. Note, make sure that the only time derivative is
∂p
∂t . Your answer should only depend on r, k, µ, p, Z, φ, cf , and t. Re-
member that µ and Z are not constant.

I will assume that φ and K are homogeneous, and an isotropic permeability, K = k. The
general conservation equation with no sources or sinks in a stationary medium and for a single
phase is,  
∂ρ ρ
φ = k∇ ∇p
∂t µ
Substituting the result from part (a) and expanding derivatives, we get,
   
∂ Mp Mp 1
0=φ − k∇ ∇p
∂t ZRT ZRT µ
Z 0 ∂p
   
M 1 M p
=φ −p 2 −k ∇ ∇p
RT Z Z ∂t RT Zµ
 p  ∂p  
p
= φ cf − k∇ ∇p
Z ∂t Zµ

(d) Upon staring at your result in part (c), you are hit by a stroke of
genius; you happen to consider the pseudopressure,
Z p
s
ψ (p) = ds,
pref µZ
where pref is a reference pressure. Rewrite the real gas pressure
equation in terms of the pseudopressure ψ.

By a direct application of the rules of differentiation we have;


∂ψ ∂ψ ∂p p ∂p
= =
∂t ∂p ∂t µZ ∂t
∂ψ p
∇ψ = ∇p = ∇p
∂p µZ
The derivative terms on the right hand side of the above expressions appear in the result from
part (c). Replacing them with the simplified derivatives yields;
φcf µ ∂ψ
− ∆ψ = 0
k ∂t

17
1.6 Practice problems
Problem 1.
Consider the new patent pending device in Figure 1.5 below. The device consists of a porous media
disc with inner boundary of radius rw and outer radius ro . The outer boundary is sealed. The
inner boundary is attached to a cylindrical vertical pipe. In the pipe is a frictionless piston of mass
M . Initially, at t = 0, the piston is at a height y = 0, and the disc and the cylinder pressures
are p = p0 . The mass is let go, and it drops gradually to an equilibrium position, y = y ∗ , as it
compresses the air in the sealed system. During this time, the pressure within the disc-cylinder
assembly evolves until the uniform equilibrium pressure, p = p∗ , is attained. Assume that the
rock is incompressible with isotropic permeability. The air is compressible and its density follows
ρ = cf p.

Figure 1.5: A schematic of a ”compressometer”.

(a) Perform a force balanc on the piston at equilibrium, t = t∗ , to derive an expression for the
pressure in the well at equilibrium, p (r = rw , t = t∗ ).
dy
(b) Perform a force balance on the cylinder for 0 ≤ t ≤ t∗ . Derive an integral expression for dt
as
a function of the cylinder pressure, p (r = rw , t).

(c) Derive an expression for the flux of air into the disc, Qw , as a function of the cylinder pressure,
p (r = rw , t). (Hint: derive an expression for Qw in terms of dydt
first.)

(d) Write the governing equation and the initial and boundary conditions for the mass balance in
the disc. All auxiliary data must be written in terms of p (r, t).

Problem 2.
Consider primary production from a reservoir that is saturated with immobile residual water
and oil. In this case, the water does not move as the oil is produced. Moreover, assume that the
change in water saturation due to changes in its density are negligible.

Let ν ∈ {o, w} denote the phase index. Then the governing equation for the conservation of
phase ν is simply,  
∂ Krν
(φρν Sν ) − ∇ Kρν ∇P = 0
∂t µν

18
Assume a constant isothermal rock and fluid compressibility, i.e.,
1 dφ
cr = ,
φ dP
and, for ν ∈ {o, w},
1 dρν
cνf = .
ρν dP
Derive the following single phase flow equation;
K∗
 
∗ ∂P
c φρo − ∇ ρo ∇P = 0
∂t µν
What are the expressions for the constants c∗ and K ∗ in terms of known quantities?

Problem 3. Conservation in non-deforming porous media


The device depicted in Figure 1.6 can be used to determine the pressure within a torus shaped
balloon as it is being filled using only direct measurements of the balloon radius R (t). The device
consists of an incompressible porous media disc of thickness H, with constant and homogeneous
porosity, φ, and permeability k. The disc has a well drilled through its center with radius rw . The
outer radius of the disc is ro . The outer edge of the disc is attached to the torus shaped balloon
and air is allowed to flow in between the disc and the balloon. Initially, the balloon is empty,
R (t = 0) = 0, and the system is at atmospheric pressure, Patm .

Figure 1.6: Cross-sectional view of annular device.

Air of constant viscosity µ and compressible density ρ (p (x, t)) is injected at a constant injec-
tion pressure pw as the outer annular balloon expands. The inventor measures the radius of the
balloon as a function of time and it turns out to be R (t).

19
(a) Derive a conservation of mass equation for air in the balloon. You may assume that the pressure
within the balloon is homogeneous and is equal to the pressure at the outer edge of the porous
media disc.

(b) Derive a conservation of mass equation for the flow through the porous media disc. You may
assume radial symmetry.

(c) Write initial and boundary conditions for the pressure in the porous media disc.

Problem 4. Conservation in deforming porous media


Consider the patent-pending device in Figure 1.7 called the extraction core-crusher. The device
consists of a homogeneous slab of rock that is saturated with oil. Two metal plates are used to
squeeze the slab in the y-direction, thereby producing the oil that was inside of the slab from the
open ends on the left and right. The ambient (outside) and initial pressure are patm .

Figure 1.7: Schematic of the extraction core-crusher device

The slab has a porosity that depends on pressure, φ (p), and a permeability that is constant,
k. The oil has a pressure dependent density ρ (p) and constant viscosity µ.

The velocity of crushing is given as in the figure and it is controlled by a hydraulic device that
squeezes the upper and lower plates together.

(a) Given that the initial position vectors for any point are x (t = 0) = x0 i + y0 j, solve for the
position of all points in the slab for any later time, x (t > 0). Use your solution to express
H (t) as a function of H0 .

20
(b) Derive the oil conservation of mass equation in terms of pressure along with appropriate bound-
ary and initial conditions. Start with an integral form and arrive at a Partial Differential
Equation in terms of pressure.

(c) Write explicit boundary and initial conditions that the pressure must satisfy.

Problem 5.
Consider an annular reservoir that is homogeneous and saturated with water. The reservoir has
a homogeneous and constant permeability, k, and porosity that depends on pressure as φ = cr P ,
where cr is the constant isothermal compressibility. The water may be considered incompressible
with constant density, ρ, and viscosity, µ.

Initially, t = 0, the reservoir has an internal radius of rw and an outer radius of ro . For t > 0,
the reservoir is being deformed radially so that the velocity of any point in the reservoir is,
dr
(t) = (rw − r (t)) e−t
dt
This situation is depicted in Figure 1.8

Figure 1.8: A radially deforming reservoir

As the reservoir is being deformed, there is no flow through its outer boundary, while the pres-
sure within the inner boundary is maintained at P0 . Inside the reservoir, the pressure will depend
on space and time; P (r (t) , t).

21
(a) Initially, a certain point within the reservoir had the coordinates r0 er + θeθ . At some later
time t > 0, where will this point be located with respect to the fixed polar coordinate system?

(b) Write the complete conservation of mass equation for water within the reservoir along with
complete boundary and initial conditions. Use polar coordinates and expand derivatives. The
result should only involve the following terms: t, r (t) , rw , ro , cr , k, µ, ρ, P (r (t) , t)

Problem 6.
Consider two oil reservoirs connected by a conduit along a fault such that fluid can flow from one
reservoir to another as shown in Figure 1.9. In this figure, p1 (t) and p2 (t), respectively, represent

Figure 1.9: Fault Block Model

average pressure in Reservoir 1 and Reservoir 2 where pressures are in psi. V1 and V2 , respectively,
represent the pore volume and of Reservoir 1 and Reservoir 2 in ft3 . Assume incompressible rock
and that fluid compressibility, denoted by c (psi−1 ), may be treated as constant, such that ρ = c.p.
At time zero, both reservoirs are at the same initial pressure, pi , in psi. For t ≥ 0, Reservoir 1 is
produced at a constant rate q ft3 /day. Define the pressure drop between the two reservoirs by

∆p(t) = p2 (t) − p1 (t).

Assume fluid does not flow from Reservoir 2 to Reservoir 1 until the pressure drop exceeds ∆b
p so
the flow rate is 



 0 if ∆p(t) ≤ ∆b p
q2,1 (t) = ,



β∆p(t) if ∆p(t) > ∆b p

where β is a constant.

22
(a) Find the time at which fluid begins to flow from Reservoir 2 to Reservoir 1.
(b) Assuming the pressure drop between the two reservoirs is constant, find the value of β.

Problem 7.

Figure 1.10: Cross-sectional view of the Ocean floor.

An urgent situation has arisen due to a recent accident (Figure 1.10); a large blob of contaim-
inated sea water is detected inside a deep sea trench. There is serious concern about how the
contaiminant will migrate into the sea floor over time. You have been tasked with providing an
engineering assessment of the situation. You decide to make some reasonable assumptions that
would allow you to devise a suitable model of the situation.

Figure 1.11: Intial condition

Consider the simplified domain Ω as depcited in Figure 1.11 below. You will develop a single
model for the square rock domain. The rock has a constant clean water supply at the top face
at a constant pressure of P0 . The right and bottom faces of the rock domain are assumed to
be sealed and no flow occurs across them. The left face between the rock and the trench can

23
experience flow. There is an infinite supply of clean and contaminated water in the trench, and
so the height of clean-contaminated contact remains constant over time (0.5L). The fluids are
consdered incompressible so that denisty and viscosity are constant and the contaminant density is
five times that of the water. Assume that the velocity of the clean and contaminated water insode
the rock is the same and obeys Darcy’s law,
K
uc = uw = − ∇P.
µw

Also you can denote the saturation of contaimnated water in the rock as Sc (x, y, t) = 1−Sw (x, y, t),
where Sw (x, y, t) is the cleant water saturation.

(a) Write a mass conservation differential equation for the contaminant inside the rock at any time
in terms of quantities introduced in the problem statement (15 points).

(b) Write a mass conservation differential equation for the clean water inside the rock at any time
in terms of quantities introduced in the problem statement (15 points).

(c) Write 2 boundary condition equations that must always be satisfied on the right wall of the
rock (10 points)

(d) Write 2 boundary condition equations that must always be satisfied on the bottom wall of the
rock (5 points)

(e) Write 2 boundary conditions equations that must always be satisfied on the top wall of the
rock (15 points)

(f) Suppose that the free water in the trench is in hydrostatic equillibrium at all times; i.e. the
pressure along the trench-rock iterface on the left must satisfy:
d
P (x = 0, y, t) = ρ (0, y, t) g, t ≥ 0
dy

Write a expression for the fluid density along the trench-rock wall as a function of depth (hint:
use an ”if”) (5 points). Derive an expression for the pressure along the trench-rock wall in
terms of quatities introduced in the problem (Hint: integrate) (10 points).

(g) Write an expression for the contaminant saturation along the left wall (Sw (x = 0, y, t) =?) (5
points)

(h) While the initial saturation of contaminated water in the rock is zero (Sc (x, y, t = 0) = 0 for
any x ∈ [0, L] and y ∈ [0, L]), the initial pressure must be determined. Assuming that initially,
the pressure in the rock must satisfy,

P (x, y, t = 0) = g [Sw (x, y, t = 0) ρw + Sc (x, y, t = 0) ρc ] ,
∂y

determine a closed form expression for the initial pressure field in the rock (15 points).

(i) Identifies the unknowns in your model and indicate the equations that you would use to solve
them (5 points).

24
Problem 8.
A recent article in the Journal of Petroleum Technlogy remarks on two sessions at the URTe
Conference in 2018. These sessions drew overflow crowds to witness findings from the Hydraulic
Fracturing Test Site project. According to actual measurements, hydraulic fracture geometry is
far more complex than ever previously modeled to date. There is a timely and important call for
the community to build more realistic models that have any hope of matching reality. As aspiring
innovators, we must first learn about the classical (and apparently wrong) oversimplified models.
In this question, we will develop such a model.

The figure depicts an idealized fracture (f) inside a matrix (m). The image is an areal slice
(neglect gravity) where a wellbore of radius rw is completed, and a fracture of aperture (width)
hf and half-length lf leads into the reservoir domain of dimensions hm by lm . Both the matrix
and fracture are assumed not to deform and are homogeneous. The fracture and matrix have very
different homogeneous and constant permeability and porosity. Assume that the well is operated
at a constant pressure Pwf all along the circumference of the wellbore-sandface perimeter. The
reservoir (fracture and matrix) is initially at a uniform pressure of P0 that is significantly larger
than the well pressure, i.e. P0 >> Pwf . Initially, the pore space is completely saturated with
natural gas of density ρ and viscosity µ. The dry gas does not condense into liquid at any pressure
in between P0 and Pwf . The density and viscosity in this pressure range however may vary with
pressure. While the right boundary of this domain on which the well is completed is considered to
have no flow across it at any time, the top, bottom, and left boundaries are assumed to remain at
the same pressure for a very long time. This is because the transient pressure wave does not have
enough time to reach the boundaries.

1. Develop a complete mathematical model that if solved, provides the pressure at every point
at any time within the reservoir; P (x, y, t) for every point (x, y) inside the rock (fracture or
matrix) and any time t ≥ 0. You may follow the steps below:

(a) Decide on an origin for an x − y coordinate system and specify it by a simple sketch.
(b) Break the rock into two subdomains (fracture and matrix parts) and denote them and
their boundary coordinates using the same coordinate system and origin decided above.
(c) Develop a differential equation for the fracture including initial and boundary conditions.
(d) Develop a differential equation for the matrix including initial and boundary conditions.
(e) Explain how the two models above are tied together.

25
2. In words, please explain how this model should change or behave in the case that the domain
is huge compared to the very small fracture aperture; i.e., assuming that hf << hm

Problem 9.
Recall the general continuity equation from physics,
Z Z Z
d
m (x, t) dx = −q (x, t) · n̂ dx + s (x, t) dx,
dt Ω(t) ∂Ω(t) Ω(t)

in a spatial domain Ω that may be moving with respect to a stationary reference frame.
Derive the mass conservation partial differential equations for two phase-flow in rock assuming that
the two fluid phases do not exchange mass. Define each quantity that you introduce as well as it
units in terms of mass M , length L, and time T ; e.g.

K: Darcy permeability L2
 

P : fluid pressure M L−1 T −2


 

φ: void porosity [1]

26

You might also like