Você está na página 1de 12

ARMA/NARMS 04-494

Sand Production and Instability Analysis in a Wellbore using a Fully


Coupled Reservoir-Geomechanics Model
J. Wang1, R. G. Wan2, A. Settari3, D. Walters4, and Y. N. Liu5
1,4

Taurus Reservoir Solutions Ltd., 2,5 Department of Civil Engineering, University of Calgary, 3 Department of
Chemical and Petroleum Engineering, University of Calgary
Copyright 2004, ARMA, American Rock Mechanics Association
This paper was prepared for presentation at Gulf Rocks 2004, the 6th North America Rock Mechanics Symposium (NARMS): Rock Mechanics Across Borders and Disciplines, held in Houston,
Texas, June 5 9, 2004.
This paper was selected for presentation by a NARMS Program Committee following review of information contained in an abstract submitted earlier by the author(s). Contents of the paper, as
presented, have not been reviewed by ARMA/NARMS and are subject to correction by the author(s). The material, as presented, does not necessarily reflect any position of NARMS, ARMA,
CARMA, SMMR, their officers, or members. Electronic reproduction, distribution, or storage of any part of this paper for commercial purposes without the written consent of ARMA is prohibited.
Permission to reproduce in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgement of where and by
whom the paper was presented.

ABSTRACT: This paper presents a fully coupled reservoir-geomechanics model with erosion mechanics to address wellbore
instability phenomena associated with sand production within the framework of mixture theory. A Representative Elementary
Volume (REV) is chosen to comprise of five phases, namely solid grains (s), fluidized solids (fs), oil fluid (f), water (w) and gas
(g). The particle transport and balance equations are written to reflect the interactions among phases in terms of mechanical
stresses and hydrodynamics. Constitutive laws (mass generation law, Darcy's law, and stress-strain relationships) are written to
describe the fundamental behaviour of sand erosion, fluid flow, and deformation of the solid skeleton respectively. Subsequently,
the resulting governing equations are solved numerically using Galerkins method with a generic nonlinear Newton-Raphson
iteration scheme. Numerical examples in a typical light oil reservoir are presented to illustrate the capabilities of the proposed
model in the absence of the gas phase. It is found that there is an intimate interaction between sand erosion activity and
deformation of the solid matrix. As erosion activity progresses, porosity increases and in turn degrades the material strength.
Strength degradation leads to an increased propensity for plastic shear failure that further magnifies the erosion activity. An
escalation of plastic shear deformations will inevitably lead to instability with the complete erosion of the sand matrix. The selfadjusted mechanism enables the model to predict both the volumetric sand production and the propagation of wormholes, and
hence instability phenomena in the wellbore.

1. INTRODUCTION
The production of formation sand has plagued the
oil and gas industry for decades because of its
adverse effects on wellbore stability and equipment,
while it has also been proven to be a most effective
way to increase well productivity. When
hydrocarbon production occurs from shallow and
geologically young (or so-called unconsolidated /
weakly consolidated) formations that have little or
no cementation to hold the sand particles together,
the interaction of fluid pressure and stresses within
the porous granular material can lead to the
mechanical failure of the formation and unwanted
mobilization of sand. It has been reported that 10%40% sand cuts normally stabilize in time to levels
less than 5% in heavy oil reservoirs [1], while an
average of 40% productivity increase was achieved
through sand management in light oil reservoirs [2].
When sand is produced from reservoir formations, it
can cause a number of problems. These include the
instability of wellbores, the erosion of pipes, the
plugging of production liners, the subsidence of

surface ground, and the need for disposal of sand in


an environmentally acceptable manner. Each year,
these issues cost the oil industry hundreds of
millions of dollars. Furthermore, sand production
and control becomes extremely crucial in offshore
operations where a very low tolerance to sand
production is allowed. Hence, it is imperative to
find an efficient computational model that has the
predictive capability to assist field operators to
understand this unique process. The ultimate goal is
to design an economical well-production strategy in
which sand production and operating costs may be
reduced to some extent with maximum hydrocarbon
productivity. It is commonly believed that the
mechanism of sand production can be attributed to
geomechanics and multi-phase or foamy oil effects.
However, modelling such a complex problem is a
challenging task since it requires multidisciplinary
physics to capture the whole range of material
response from sand flow initiation to fluidization.
In this paper, sand production is treated as an
erosion process by which a weakly consolidated
sand matrix is disaggregated near perforations of a

wellbore due to a combination of stress changes and


multiphase flow. A fully coupled reservoirgeomechanics mathematical model is presented to
account for the effects of multiphase flow and
geomechanics as well as their interaction in a
consistent manner. Numerical solutions, restricted
to a typical light oil reservoir without the influence
of the gas phase, are sought to examine the basic
capabilities of this model. As the wellbore pressure
is lower than reservoir pressure, the erosion process
begins as a result of the degradation of the sand
matrix strength and the drag force imposed by fluid
pressure gradient. The plastic yielding zones
develop due to the material degradation (erosion)
and stress re-distribution, while the wormholes or
cavities form and propagate in terms of the
increasing porosity values. The volumetric oil and
sand productions are also calculated as a function of
time, stresses, and hydrocarbon flow rate.
2. COUPLED MULTIPHASE FLOW
GEOMECHANICS FORMULATION

AND

2.1. Mass balance equations


The single-phase formulation describing sand
production in a deforming sand matrix was derived
in a series of publications [3, 4]. It has been shown
to be a promising method for modeling sand
production in terms of matching numerical
calculations with lab test data, both in heavy and
light oil conditions [5, 6, 7]. In this paper, an
extension to multiphase sand production model is
presented within the same framework of mixture
theory, i.e., a coupled black-oil/geomechanics sand
production model with erosion mechanics is
proposed to further account for the effects of
multiphase flow of three components (gas, water,
oil) and their interaction with geomechanics. The
mass balance equation used in formulating the sand
production problem is typically written as

+ (u& ) = m&
t

(1)

where state variables , u& are the density and the


absolute velocity respectively, and m& is the source
or sink term to account for the local rate of solid
loss or gain per unit volume due to erosion.
The fluid/gas saturated sand body is idealized as a
Representative Elementary Volume (REV) which
comprises of five phases, namely solid grains (s),
fluidized solids (fs), fluid (f), water (w) and gas (g)

as shown in Figure 1. In reality, the individual


distribution varies discontinuously over space.
However, an averaging procedure in the spirit of
mixture theory is used to homogenize each
constituent over the REV volume V such that these
individuals are substituted with continuous ones that
fill the whole volume. Each phase discontinuity in
the REV is represented in terms of its own volume
fraction, i.e. saturation and porosity.

wellbore
sand, oil,
gas
cavity or
wormhole

free & disolved gas


gas (dg+fg)
fluidized solids (g) M , , dV
g g
g
fluid
fluid
solids
(f) Mf , f , dVf

REV

fluidized solid
(fs) Mfs , fs , dVfs
solid
(s) Ms , s , dVs

dVv
dV

Phase diagram

Fig. 1 Phase components of a REV

For solid phase (s), the density of the solid phase


averaged out over a REV of volume dV can be
written as the homogenized solid density (1-)s ,
V
where porosity = dV
dV , and s is the density of the
solid phase. The mass conservation requires that

[(1 ) s ]
+ [(1 ) s u& s ] = m&
t

(2)

where u& s is the absolute velocity of the solid phase


boundary, and the negative sign of the right hand
side refers to a solid loss due to erosion since m& is
chosen to be the local rate of solid gain per unit
volume as seen from the fluidized solid phase.
Similarly, for the fluidized solid phase (fs), the mass
balance equation can be written, i.e.

[S fs fs ]
t

+ [S fs fs u& fs ] = m&

(3)

where the fluidized solid saturation at reservoir


[dV ]RC
condition (RC) is S fs = [dVVfs ]RC
, u& fs is the absolute
velocity of the fluidized solid phase, and fs is the
density of the fluidized solid phase.
The basic assumptions for flow of oil, water and gas
phases follow those used in the classical black-oil
model [8]. The oil phase (o) continuity equation can
be derived at stock tank condition (STC), i.e.
[S o o / Bo ]
+ [S o o u& o / Bo ] = 0
t

(4)

where o = fluid density at stock tank condition,

So =

[Vo ]RC
[VV ]RC

= oil saturation in reservoir condition


[V +V ]
(RC), Bo = o[Vo ]dgSTCRC = the formation volume factor,
and u& o = the absolute velocity of the oil phase.
Furthermore, the averaged density of gas can be
divided into two components: free gas S g g / Bg
[V ]
and dissolved gas gSo , where S g = [VVg ]RCRC ,
[V ]
Bg = [Vgg ] RC , g = BRos g , g = the gas density at
STC
[V ]STC
. Hence, the
stock tank condition, and Rs = [Vdgo ]STC

mass balance for the gas phase is written, i.e.


[S g g / Bg + Rs So g / Bo ]

t
+ [S g g / Bg u& g + Rs So g / Bo u& o ] = 0

(5)

Since the water is assumed not to partition in either


the hydrocarbon liquid or the gas phase, the mass
balance for the water phase is given as
[S w w / Bw ]
+ [S w w u& w / Bw ] = 0
t
where S w =

[V w ]RC
[VV ]RC

, Bw =

[Vw ]RC
[Vo ]STC

(6)

can be related to a

function involving water phase pressures.


In the above, the velocities u& o , u& g and u& w are
defined somewhat differently from what is
customary done in the multiphase flow literature.
They are interstitial velocities, based on an
assumption that the flow area Aj for the any phase j
is equal to the total pore (void) area AV times the
phase saturation Sj. Therefore, the absolute velocity
u& j is related to Darcy velocity v j (see Eq. (9) that
follow in the next section).
2.2. Equilibrium equation for the solid matrix
The interaction between the mechanical behaviour
of a deforming solid matrix and fluid dynamics
must be incorporated into the governing equations
in order to describe the coupling effects. The
volume-weighted solid velocity u& s provides the
linkage between the fluid and geomechanical
aspects of the problem. The latter involves a
deforming sand skeleton under an effective stress
field eff and the volume-averaged pore mixture
pressure Pm, which must satisfy momentum
balance, i.e.

( eff Pm 1) + b = 0

(7)

where b are body forces per unit volume, and is


a parameter accounting for the compressibility of
the sand grains. The sign convention adopted is that
negative stresses are compressive and fluid
pressures are always positive. The Kronecker delta
tensor is given by 1 such that 1ij = ij . The
averaged mixture pressure can be defined as
Pm = So Po + S g Pg + S w Pw

(8)

2.3. Discharge for each phase


In anticipation for the description of fluid flow
through a porous medium, a volume averaged
discharge velocity v j (j= o, w, g) of each fluid
phase relative to the solid matrix (Darcy velocity) is
defined as

v j = S j (u& j u& s )

(9)

Both the detachment and fluidization of solid


particles are a dynamic process that is complex in
nature. It is a future research task to define the
interaction between fluidized particle and fluid at a
micro/macro level. However, the discharge of
fluidized solid phase can be related to the average
velocity of mixture, i.e.

v fs = S fsu& fs = S fs ( v m u& s )

(10)

where the average velocity of mixture is

v m = So v o + S g v g + S w v w

(11)

Eqs.(2-6) represent local mass balance equations for


each individual phase. Successively combining
these equations with Eqs.(9-10), the following five
governing equations are obtained for each phase,
i.e.

m&
+ [(1 ) u& s ] =
t

[ ( S fs 1)]
t

+ [S fs v m + (1 + S fs ) u& s ] = 0

(12)

(13)

v
S u& S
. o + o s + o = 0
Bo t Bo
Bo

(14)

v
S u& S
. w + w s + w = 0
Bw t Bw
Bw

(15)


S u&
R S u&
v g / Bg + g s + Rs v o / Bo + s o s
Bg
Bo

S g
R S
+ s o +
=0
Bg
t Bo

(16)

2.4. Constitutive laws


Eqs.(12-16) must be supplemented with constitutive
laws describing sand particle erosion, fluid flow,
and deformation of the sand matrix. It is commonly
believed that the driving force causing the solid
detachment from the sand matrix is due to
hydrodynamics and geomechanics. Based on
phenomenology, a possible functional form of mass
generation can be obtained from the inverse of
filtration theory as proposed in refs. [9,10], i.e.

m&

= (1 ) S fs v m

if v m v crm

=0

if v m < v crm

(17)

where v crm is the critical average velocity of mixture


below which no sand production occurs. The
erosion coefficient provides a length scale that
can be linked to the accumulated plastic strains p
through the following relationship, i.e.

= ( p ) = 0 +

p
p / max
p
+ p / max

(18)

where and are constants to be determined,


p
corresponds to the
while 0 is a constant, and max
maximum plastic shear strain calculated for the
entire domain. Eq.(18) gives a hyperbolic variation
of with respect to normalized plastic shear strain
p
p / max
, i.e. with increasing plastic strains,
becomes larger which in turn increases erosion
activity as implied in Eq.(17), see Figure 2.

As for describing fluid flow, semi-empirical Darcy's


law is used to establish the relation between
pressure gradient Pj and volumetric fluid mixture

vj =

Pj

(19)

where j= o, w, g, m, k is the effective permeability


tensor that can be related to porosity via the
Carman-Kozeny equation or its variant, i.e.

= 0 +

p
p / max
p
p
+ / max

p
p / max

Fig. 2 Relation between erosion and plastic shear strain

k = k0

0
3
1 or k = k0 exp A
1
2
(1 )
1

(20)

where k 0 and 0 are constants and A is a fitting


parameter. Turning to solid skeleton deformations, a
more adequate constitutive law based on plasticity
and incorporating stress dilatancy aspects must be
used, considering that sand behaviour is mainly
dissipative and dominated by grain slippage,
rearrangement, dilation and destructuration. A yield
function F ( ) based on Mohr-Coulomb is
considered adequate, while a plastic potential
function G must be introduced to calculate plastic
strains. The flow rule basically defines the plastic
strain increment vector as the normal to the plastic
potential function G and its magnitude determined
from the plastic multiplier , i.e.
G
eff
d 0 if F ( ) = 0 and dF = 0
d = 0 if F ( ) = 0 and dF < 0

d p = d

(21)

In order to describe the internal damage due to the


degradation of the porous medium as the erosion
proceeds, it is assumed that the material properties
such as cohesion C and friction angle drop
linearly with porosity , i.e.

C = C0

flux per unit area v j . Thus,

max

1
1
and = 0
1 0
1 0

(22)

where C 0 and 0 are constants. Such a simple


damage law enables this model to account for the
effect of the degradation of the porous medium due
to the erosion. More precisely, the plastic
deformation of sand matrix increases the erosion

potential according to Eq.(18). In return, the erosion


process also weakens the sand matrix through
degradation of its strength properties, see Eq.(22).
In order to complete the derivation of governing
equations, we have to define the capillary pressure
Pc relationship. The most practical method is to use
an empirical correlation relating the capillary
pressure and phase saturations [8], i.e.
Pcow = P0 Pw = f ( S o , S w )
Pcog = P0 Pw = f ( So , S g )

(23)

In conclusion, we have eight equations for solving


eight
field
unknowns,
namely,
, S fs , Pj ( j = o, g , w) and u i (i = 1, 2, 3) in the
three-dimensional case.
3. STABILIZED FINITE ELEMENT
SOLUTIONS

Wn +1 ( Vn +1 ) = H n +1 ( Vn )

(25)

in which W and H are functionals which originate


from Eqs.(12-16) and subscripts n and n+1 refer to
time stations t n and tn +1 respectively. Eq.(25)
represents the standard non-linear matricial
equations that can be solved via iterative schemes
such as the Newton-Raphson method. If superscript
k denotes the iteration number during successive
attempts to final solution, then expanding Eq.(25)
using the Taylors series leads to
k

Although the writing of the governing equations is


rather straightforward, both their finite element
discretization and solution are challenging due to
the nature of the equations and field variables.
Numerical instability arises in terms of node-tonode oscillations. Over the past several years, the
authors developed a generic numerical stabilization
scheme - an optimized local mean technique. By
enriching main field variables with high gradient
terms, sharp non-local changes can be captured in
the computations to ensure stable solutions. Then,
the enriched field variables enter into the governing
equations of physics by way of averaging of the
field values in the neighbourhood of a continuum
point, see details in [11]. Thereafter, the finite
element discretization of the modified governing
equations is ready to be expressed in terms of
variables V, i.e. the nodal displacement
ui (i = 1, 2, 3) , phase pressure P j ( j = o, g , w) ,
porosity , and fluidized sand saturation S fs .
V ( x, t ) = N k ( x ) V (t )

weighting functions equal to interpolation


functions) over the entire domain to above
governing equations in turn together with
discretizing time derivatives by standard finite
difference formula and also linearizing time
variables, a system of five non-linear equations is
obtained with its generic form, i.e.

(24)

where V stands for S fsp , p , p jp , uip , and N p are


respectively fluidized solid saturation, porosity,
fluid pressure, displacement, and interpolation
function at node p, for p=1 to nh , the total number
of nodes. It is again recalled that Einstein index
notation is used with repeated indices implying
summation and the index p is dummy. Applying
Galerkins method of weighted residual (with

Wnk+1 ( Vnk+1 ) +

W
Vnk+1 = H n +1 ( Vnk )
V n +1

(26)

Hence, the increment of vector V at the end of


iteration k is

[ ] [H

Vnk+1 = J kn +1

n +1

( Vnk ) Wnk+1 ( Vnk+1 )

(27)

k
in which J n1 is the Jacobian of the linearized
system, i.e.
k

k
n +1

W
=
V n +1

(28)

Successive iterations are performed until the


convergence criteria are satisfied, i.e.
Vnk++11 Vnk+1 <

(29)

where Vnk++11 = Vnk+1 + Vnk+1 and is a small value.


Hence, the incremental form of the equations to be
solved at the element level emerges as
[A1]
[B1]

[C1]

[D1]

[A2]
[B2]
[C 2]
[D 2]

[A3]
[B3]
[C 3]
[D3]

[A4] k S kfsn+1
[B 4] nk+1
[C 4] p kn+1
[D 4] n+1 u kn+1

= X ( S kfsn+1 , nk+1 , p kn +1 , u kn +1 )

(30)

where [X] a vector containing known variables at


previous iteration, and [A1], [A2],... [D2], [D3] are

sub-matrices pertinent to fluid, solid, fluidized


solid, and stress-deformation properties [12]. The
procedures of Newton-Raphson algorithm are listed
in Table 1. From a practical point view, we have to
address properly the various coupling strategies, i.e.
decoupling, explicit, and implicit coupling
techniques before proceeding with the fully coupled
reservoir/geomechanics simulation [13].
Table 1. Procedures for Newton-Raphson scheme
1. Set the initial value k=0 and initial values for each variable
k

2. Calculate the Jacobian matrix J n +1 according to Eq.(28)


3. Calculate the right hand side X in Eq.(30)
4. Solve Eq.(30)
5.Check for convergence
IF: Eq.(29) is satisfied THEN
Go to next time step
ELSE
Go to : 2 with new trial value for each variable and k=k+1
ENDIF

4. NUMERICAL EXAMPLES
In the following simulation, a numerical example of
a light oil reservoir in North Sea is examined under
hydrodynamics and geomechanics, while examples
in heavy oil reservoirs can be found in a series of
publications [5, 6, 7]. In this paper, no gas phase
effect is presented, given the space restriction.
0.5

0.4

are chosen to be 0.001 and 0.25 respectively. The


simulation is conducted as follows. First, the initial
state of the reservoir is computed based on an oil
saturation pressure of 27.6 MPa and an external
stress of 42 MPa is imposed on both wellbore and
outer boundaries. Then, the stress around wellbore
is changed to a reservoir pressure of 27.6 MPa to
simulate the open-hole completion. Finally, a 3
MPa drawdown is applied at three perforations (P1,
P2, and P3) as shown in Figure 3.
The length of each perforation is 0.25 m with a
0.012m diameter for P2, and a 0.006m diameter for
both P1 and P3. These, in fact, refer to eight
perforations for the full well configuration. The
initial porosity and erosion coefficient in the
perforations are set to 0.6 and 3 m-1 respectively to
account for the disturbance caused by the
perforation process, while they are set to 0.25 and 2
m-1 in the remainder part of the reservoir formation.
Finally, the entire finite element grid is comprised
of 3840 nodes and 3705 4-nodes elements and the
time step size used in the analysis is 0.005 day for a
total time span of 5 days investigated. Table 2
shows the material properties (fluid and
geomechanics) used in the simulation.
Table 2. Model parameters
0 = 2 or 3 m-1
s = 2.7 g/cm3
K0x = 0.5 Darcy K0y = 0.1Darcy
C0 = 6 MPa
E = 2 GPa

ext = 42MPa
0 = 30
=0.008
=0.1

o = 0.8 g/cm3
= 5 cp
= 0.25
P0= 27.6 MPa

For the purpose of clarity of illustration, the figures


are plotted in the vicinity of the wellbore, within the
first 1 m, 2 m and 5m as indicated in XY axes
respectively.

0.3

extends to 5 m
Fig. 3. Mesh layout near wellbore showing perforations.

4.1. Deformations and yielding after open-hole


completion and perforations
In order to examine the wellbore instability and
sand production, it is essential to understand the
open-hole completion and perforation process. The
process is simulated by lowering the initial stresses
42 MPa at inner holes to the initial reservoir
pressure and the outer ones are kept to initial stress
conditions after reservoir initialization.

Figure 3 shows a close-up of the finite element


mesh representing one quarter of a section of a
vertical well of inner radius r0 = 0.1 m with the
outer boundary of the well extending to 5 m. The
initial fluidized sand saturation Sfso and porosity 0

It is noted that a plastic zone is developed as shown


in Figure 4. This is due to the stress re-distribution
around wellbore and the existence of a weakened
zone in the perforations (0=0.6) during the drilling
process. It is critical to capture the developed plastic
zones due to drilling and perforation, since the

0.2

0.1

P3
P2

perforations
0

P1

0.1

0.2

x(m)

0.3

0.4

0.5

time=0.3days

0.14
0.13
0.12
0.11
0.11
0.10
0.09
0.08
0.07
0.06
0.05
0.04
0.03
0.02
0.01

0.75

y(m)

erosion coefficient is linked to plastic shear strain


as defined in Eq.(18) - the larger the plastic shear
strains are, the more intensive the erosion activity
is. This enables the simulator to automatically
capture the disturbance caused by open-hole
completion and perforation in terms of the initial
values of erosion coefficient and porosity around
wellbore and perforations at the beginning of the
drawdown.

0.5

1
0.25

P1

0.75
0

P2
P3

0.25

0.5

0.75

x(m)

y(m)

Fig. 5 Fluidized sand saturation profile at time t=0.3 day.


0.5
2

Plastic yielded zones


time=0.6days

0.25

0.14
0.13
0.12
0.11
0.11
0.10
0.09
0.08
0.07
0.06
0.05
0.04
0.03
0.02
0.01

1.5

P1

P2
P3

0.25

0.5

x(m)

0.75

y(m)

Fig. 4 Plastic yielded zones developed after open-hole


completion and perforations (before drawdown).
0.5

0.5

1.5

x(m)

Fig. 6 Fluidized sand saturation profile at time t=0.6 day.


5

time=2days

0.14
0.13
0.12
0.11
0.11
0.10
0.09
0.08
0.07
0.06
0.05
0.04
0.03
0.02
0.01

y(m)

4.2. Evolution of fluidized sand saturation


From this section on, we look at the field variable
profiles due to drawdown. Figures 5-7 illustrate the
spatial distribution of the fluidized sand saturation
Sfs at four different times t=0.3 day, 0.6 day, 2 days
and 5 days after drawdown. It is noticed that a sharp
rise in fluidized sand saturation develops in the
region near the perforations P1 and P2 with the
remaining part of the well being at near initial
values of Sfso. The amplification factor for fluidized
sand saturation near the perforation, defined as the
current saturation value over the initial one, is about
70 times at location P1 for time t=0.3 day, 110
times at location P2 for time t=0.6 day, and 140
times at location P3 for time t=5 days respectively.
These numbers indicate that there is a dramatic
increase in the creation of fluidized sand
corresponding to sand production. In general, an
increase in fluidized sand saturation is governed by
the relative rates at which volume of fluidized sand
Vfs and void volume VV are changing, since Sfs =
Vfs/VV. This sharp change is due to the physics of
the problem described as follows. Initially, erosion
preferentially occurs in the x-direction near

x(m)

Fig. 7 Fluidized sand saturation profile at time t=2 days.

Figure 8 shows a decreased fluidized sand


saturation profile, which indicates a decline in
erosion activity because there is no material left for
the erosion around wellbore.

of wellbore and perforations, very high fluid fluxes


prevail, which in turn give way to high fluidized
2

12.00
11.29
10.57
9.86
9.14
8.43
7.71
7.00
6.29
5.57
4.86
4.14
3.43
2.71
2.00

time=0.3days

1.5

y(m)

perforation P1 since the horizontal permeability is


five times greater than the vertical one. As most of
the sand particles are mobilized to produce a very
loose matrix, further erosion takes place in regions
where more sand particles are available.

0.5
4

time=5days

0.14
0.13
0.12
0.11
0.11
0.10
0.09
0.08
0.07
0.06
0.05
0.04
0.03
0.02
0.01

y(m)

x(m)

Fig. 8 Fluidized sand saturation profile at time t= 5 days.

4.3. Evolution of erosion coefficient and cavity


propagation
As defined in Eq.(17), the erosion coefficient is a
function of plastic shear strain. This indicates that
most erosion activity is confined and intensified in
only plastic shearing regions. The larger the plastic
shear is, the more intensive the erosion is. In other
words, the erosion activity aligns itself with the
plastic yielded zones where plastic shearing of the
material is most prevalent. Figures 9-11 show the
distribution of erosion coefficient with time
around the wellbore. The erosion activity is most
intense around the wellbore and perforations at the
very beginning, and then propagates further inside
the perforations where the sand matrix has a weak
material strength (initial porosity 0.6), and in the xdirection where the pore pressure depletion is the
fastest due to high permeability in x-direction
initially. This is due to increasing erosion activity
taking place as porosity increases and ultimately
degrades the material strength. These will be
discussed in later sections.
Figure 12 shows the initiation of erosion at the
perforations at time t=0.3 day. In fact, at the edges

y(m)

1.5

12.00
11.29
10.57
9.86
9.14
8.43
7.71
7.00
6.29
5.57
4.86
4.14
3.43
2.71
2.00

time=0.6days

0.5

0.5

x(m)

1.5

Fig. 10 Erosion coefficient distribution at time t=0.6 day.


2

12.00
11.29
10.57
9.86
9.14
8.43
7.71
7.00
6.29
5.57
4.86
4.14
3.43
2.71
2.00

time=5days

1.5

y(m)

x(m)

1.5

0.5

Fig. 9 Erosion coefficient distribution at time t=0.3 days.

0.5

0.5

x(m)

1.5

Fig. 11 Erosion coefficient distribution at time t=5 days.

sand mass fluxes as dictated by the erosion law, see


Eq.(16). However, the maximum erosion activity

x(m)

0.5

0.5

0.4

time=0.3days

0.3

y(m)

0.77
0.73
0.70
0.66
0.63
0.59
0.56
0.53
0.49
0.46
0.42
0.39
0.35
0.32
0.28

0.75

0.25

0.2

0.25

0.5

x(m)

0.75

0.1

P1
P2

Fig. 12 Porosity profile at time t=0.3 day.


0
2

0.77
0.73
0.70
0.66
0.63
0.59
0.56
0.53
0.49
0.46
0.42
0.39
0.35
0.32
0.28

1.5

0.5

P3

0.1

0.2

x(m)

0.3

0.4

0.5

0.4

0.5

Fig. 15 Fluid flux profile at time t=0.3 days.


0.5

0.4

time=0.6days

0.3

y(m)

Porosity

time=0.6days

y(m)

0.77
0.73
0.70
0.66
0.63
0.59
0.56
0.53
0.49
0.46
0.42
0.39
0.35
0.32
0.28

Porosity

time=0.3days

Porosity

time=5days

Fig. 14 Porosity profile at time t=5 days.

y(m)

y(m)

does not start simultaneously at all perforations as


shown in Figure 12. In fact, the most intensive
erosion activity follows geomechanically yielded
zones and a preferential direction of high flux, i.e.
x-direction. Figure 13 shows the coalescence of
eroded zones around perforations P1 and P2 into a
ring of loose sand of about 0.5 m in radius. The
porosity values approach 0.77 and physically
correspond to the formation of a cavity and
mechanical failure of the wellbore. Figure 14 shows
a snapshot of the fully developed zone of high
porosity that is initiated at the perforations, and
which localizes along the plastic yielded zones and
high flux regions.

0.2

0.5

x(m)

1.5

2
0.1

Fig. 13 Porosity profile at time t=0.6 day.

4.4. Fluid flux and pressure distribution


As the cavity enlarges, the permeability of the
reservoir increases since it is a function of porosity
in Eq.(19). The gradually increased permeability
enhances the well productivity. It is expected that

P1
P2

P3

0.1

0.2

x(m)

0.3

Fig. 16 Fluid flux profile at time t=0.6 days.

the high fluid flux dominates in three perforations in


Figure 15 at the beginning of drawdown. Then, the

direction of large fluid fluxes shows a bias towards


high porosity regions as shown in Figure 16, i.e.
mostly x-direction in anisotropic permeability case.
It is also worth to mention that the erosion process
increases the fluid flux by degrading the sand
matrix where more regions progressively yield
plastically due to the high fluid flux and stress
redistribution. Figure 17 shows an increased flux
region around the wellbore at time t=5 days.
0.5

4.5. Displacements and stresses


In this section, we look at the plastic shear strain
and stresses distribution in the well. The pressure
induced drag forces develop excessive plastic shear
strains around perforations in both x- and ydirection (maximum value is about 9% after 5 days
in Figure 19). It is also noted that the material
strength parameters, i.e. cohesion C and friction
angle follow the same distribution as that of
porosity with time since they are defined as a linear
function of porosity in Eq.(22).
2

0.4

time=5days

0.090
0.086
0.081
0.077
0.073
0.069
0.064
0.060
0.056
0.051
0.047
0.043
0.039
0.034
0.030
0.026
0.022
0.017
0.013
0.009
0.004
0.003
0.001
0.000
0.000

time=5days

0.3

y(m)

1.5

y(m)

0.2

0.1

P1
P2

0.5
P3

0.1

0.2

0.3

x(m)

0.4

0.5

Fig. 17 Fluid flux profile at time t=5 days.

0.5

x(m)

1.5

Fig. 19 Plastic shear strain distribution at time t=5 days.


5

(Pa)
-7.00E+06
-7.53E+06
-8.05E+06
-8.58E+06
-9.11E+06
-9.63E+06
-1.02E+07
-1.07E+07
-1.12E+07
-1.17E+07
-1.23E+07
-1.28E+07
-1.33E+07
-1.38E+07
-1.44E+07
-1.49E+07
-1.54E+07
-1.59E+07
-1.65E+07
-1.70E+07

4
time= 5 days

y(m)

Due to the initial anisotropic permeability


conditions, the dissipation of fluid pressures around
the well also occurs in regions of high
permeabilities, i.e. x-direction. As sand is being
produced, the fluid pressure slowly depletes more
from initial values of 27.6 MPa on the outside
boundary to 24.5 MPa than at perforations P1, P2,
and P3 around the wellbore, as shown in Figure 18.

5
(Pa)
2.74E+07
2.72E+07
2.70E+07
2.69E+07
2.67E+07
2.65E+07
2.63E+07
2.61E+07
2.59E+07
2.57E+07
2.55E+07
2.54E+07
2.52E+07
2.50E+07
2.48E+07

Time=5days

y(m)

x(m)

Fig. 18 Pore pressure distribution at time t=5 days.

x(m)

Fig. 20 Effective stress xx at time t=5 days.

Considering the wellbore stability, it is very


important to look at the stress distribution after sand
production. Figures 20-22 show the distribution of
effective stresses xx, yy, xy at 5 days after
drawdown. Due to fluid pressure reduction through
three perforations, drag forces are imposed upon
three perforations, causing a reduced stress xx in P3

(Pa)
-7.00E+06
-7.53E+06
-8.05E+06
-8.58E+06
-9.11E+06
-9.63E+06
-1.02E+07
-1.07E+07
-1.12E+07
-1.17E+07
-1.23E+07
-1.28E+07
-1.33E+07
-1.38E+07
-1.44E+07
-1.49E+07
-1.54E+07
-1.59E+07
-1.65E+07
-1.70E+07

4
time=5 days

y(m)

x(m)

Figure 23 gives both the oil and sand rates over the
time of fluid drawdown. We observe that the sand
production rate rapidly increases in an initial phase
to reach a peak value in approximately 0.5 day.
During this time period, the oil rate gradually
increases as well. Then, this phase is followed by a
decline in sand production rate corresponding to the
decrease in availability of sand grains. However, the
oil rate continues to increase given the enhancement
in permeability of the reservoir induced by sand
production. This trend is also observed in oilwells
under sand production.

12000

1200

10000

1000

8000

800

6000

600

oil rate
sand rate

4000
2000

time=5 days

y(m)

x(m)

25000

3000

20000

2500
2000

15000

oil rate
sand rate

10000

1500
1000

5000

500

0
0

Fig. 23 Oil and sand rate history at anisotropic permeability


conditions.

oil rate (kg/day/m)

(Pa)
3.00E+06
2.84E+06
2.69E+06
2.53E+06
2.38E+06
2.22E+06
2.07E+06
1.91E+06
1.76E+06
1.60E+06
1.45E+06
1.29E+06
1.14E+06
9.82E+05
8.26E+05
6.71E+05
5.16E+05
3.61E+05
2.05E+05
5.00E+04

time (days)

200
0

400

Fig. 21 Effective stress yy at time t=5 days.

(31)

sand rate (kg/day/m)

sand rate (kg/day/m)

qoil = v f dS ; qsand = S fs v f dS

oil rate (kg/day/m)

whereas an increased stress xx around P1 in Figure


20. Also, the stress yy is reduced in P1 and
increased around P3, as shown in Figure 21. Figure
22 shows the tangential stress profile distribution.
The high stress values indicate a highly sheared
zone. Depending on the re-distribution of pore
pressure and stress during erosion, the high shear
stress zone shifts and grows, which in turn causes
the evolution of plastic shear yielded zones.

time (days)

Fig. 22 Effective tangential stress xy at time t=5 days.

Fig. 24 Oil and sand rate history at isotropic permeability


conditions.

4.6. Volumetric sand production and oil rates


In the previous sections, detailed spatial
distributions of governing field variables with time
were discussed and the analysis revealed local
phenomena during sand production. From an
engineering point of view, we would be interested
in examining the total oil and volumetric sand
production rates as integrated over the total
perforation area S (P1, P2, and P3) of the wellbore.
Hence,

As a comparison, an initial isotropic permeability


case is also computed with kx0=ky0=0.5 Darcies. As
expected, more sand and higher oil rates are
obtained as larger initial reservoir permeability
prevails in y-direction, see Figure 24. The same
peak value of fluidized sand saturation is calculated,
but a smoother decline curve of sand rate is
obtained in isotropic case, since there is no erosion
lag due to anisotropic permeability conditions.

Journal of Canadian Petroleum Technology. 41:4, 46


52.

5. CONCLUSIONS
A fully coupled reservoir/geomechanics numerical
model is presented based on an extension of a
theoretical and numerical model that the authors
have developed in the past to address sand
production as an erosion problem coupled with
hydro- and geo-mechanical effects. This is done
within the framework of mixture theory in which
mechanics and transport equations are written for
each of the concerned phases, i.e. solid, fluid (oil,
water), gas, and fluidized solid.
Leaving aside gas-related issues, it is found that
sand production is a function of stress, time, and
fluid rate. Sand erosion activity is strongly linked to
geomechanics and there is an intimate interaction
between sand erosion activity and deformation of
the solid matrix. As the erosion activity progresses,
porosity increases and in turn degrades the material
strength. Strength degradation leads to an increased
propensity for plastic shear failure that further
magnifies the erosion activity. An escalation of
plastic shear deformations will inevitably lead to
wellbore instability with the complete erosion of the
sand matrix. The self-adjusted mechanism enables
the model to predict both the volumetric sand
production and the propagation of wormholes.
The multiphase results including gas phase will be
presented in a forthcoming paper. The proposed
model can be used for wellbore stability analysis
and design in open-hole completions, perforation
pattern design, as well as volumetric sand prediction
at different pumping strategies in terms of
optimization of the hydrocarbon production.
6. ACKNOWLEDGEMENTS
The authors wish to express their sincere gratitude
for funding provided by Alberta Ingenuity Fund
(AIF) and the National Science and Engineering
Research Council of Canada (NSERC).
REFERENCES
1.

Tremblay, B., G. Sedgwick, and D. Vu. 1999. CT


imaging of wormhole growth under solution gas
drive. SPE Reservoir Journal. 2: 1, 3745.

2.

Papamichos, E. and E. M. Malmanger. 2001. A sand


erosion model for volumetric sand predictions in a
north sea reservoir. SPE Reservoir Evaluation and
Engineering. 4450.

3.

Wan, R.G. and J. Wang. 2002. Modelling sand


production within a continuum mechanics framework.

4.

Wan, R.G. and J. Wang 2004. Analysis of sand


production in unconsolidated oil sand using a coupled
erosional-stress-deformation model. Journal of
Canadian Petroleum Technology. 43:2, 4753.

5.

Wan, R.G. and J. Wang: 2002. A Coupled StressDeformation Model for Sand Production using
Streamline Upwind Finite Elements. In Proceedings of
the Eighth International Symposium on Numerical
Models in Geomechanics NUMOG VIII, Rome, Italy,
10-12 April, 2002, eds. Pande & Pietruszczak, 301
309. A. A. Balkema, Rotterdam. ISBN 90 5809 359 X

6.

Wan, R.G. and J. Wang. 2004. Modelling of sand


production and wormhole propagation in an oil
saturated sand pack using stabilized finite element
methods. Journal of Canadian Petroleum Technology.
43:4, 4653.

7.

Wan, R. G. and J. Wang. 2003. Modeling Sand


Production and Erosion Growth under Combined Axial
and Radial Flow. SPE International Thermal
Operations and Heavy Oil Symposium and
International Horizontal Well Technology Conference
SPE 80139. Calgary, Canada, 47 November 2002.

8.

Aziz, K, and A. Settari. 1979. Petroleum reservoir


simulation. London. Elservier Applied Sci.

9.

Vardoulakis, M. Stavropoulou and P. Papanastasiou.


1996. Hydromechanical aspects of the sand production
problem. Transport in Porous Media. 22, 225-244.

10. M. Stavropoulou, P. Papanastasiou and I. Vardoulakis.


1998. Coupled wellbore erosion and stability analysis.
Int. J. Numer. Anal. Methods Geomech. 22, 749-769
11. Wang, J. and R.G. Wan. 2004. Computation of Sand
Fluidization Phenomena using Stabilized Finite
Elements, Finite Elements in Analysis and Design (in
press).
12. Wang, J. 2003. Mathematical and numerical modeling
of sand production as a coupled geomechanicshydrodynamics problem. Calgary. (PH. D. dissertation)
13. Settari, A. and D. A. Walters. 2001. Advances in
coupled geomechanical and reservoir modeling with
applications to reservoir compaction. SPE Journal. 9:
334342.

Você também pode gostar