You are on page 1of 44

Well Performance

Two interrelated components are key to understanding flow regimes the potential for injection into the
reservoir (reservoir performance) and the restrictions imposed by the completion and wellbore system
hydraulics.

Reservoir Performance
Designing a successful injector begins with careful consideration of the target formations themselves. Consider
the simplest flow situation a non-fractured reservoir, single phase, injecting/producing under pseudo-steady-
state, radial flow conditions. Darcys Law indicates that:

(1)

where:
q = flow rate (STB/D or Mscf/D at 14.7 psia and 60F, 1Mscf/D=1000 scf/D), q is
negative for injection
k = permeability (md)
h = net thickness (feet)
pwf = flowing bottomhole pressure (psi)
= static (average) reservoir pressure (psi)

pi = initial pressure (often assumed to be equal to (psi))


Bo = oil formation volume factor, which converts reservoir barrels to stock tank
barrels (RB/STB)
re = drainage radius (feet)
rw = wellbore radius (feet)
s = skin (dimensionless)
D = turbulent flow factor
= viscosity (cP)
= average compressibility factor (dimensionless)
= average temperature (R, R = F + 459.67)

Consider an example. You are planning an injection operation in an aquifer that is (or will be) into a twenty-foot
thick zone with a permeability of 1000 md. The average reservoir pressure is 2000 psi. Suppose that the
bottomhole pressure during injection is 2100 psi, the viscosity is 1 cP, the formation volume factor is 1.0, the
-5 -1
porosity is 20%, the total compressibility is 1.0 x 10 psi , and the well spacing is 20 acres. The drainage radius
is usually estimated from the well spacing. You can inscribe a circle of radius 467 feet into a square that has an
2 5 2
area of 20 acres. One acre is 43,560 ft . Twenty acres is 8.712 x 10 ft . The inscribed circle will have a radius

of . Assume that there is no skin (damage) or turbulence and that the well is drilled to 6
inches in diameter. What sort of flow rates should you anticipate?
Why do you need to know these flow rate equations? It is essential to be able to anticipate flow rates under
various bottomhole conditions that can arise during the injection operations. Looking at these rates helps to
determine if the well has been fractured, the degree of damage, etc.

Estimating rates is complicated by the fact that injection is transient (is not constant) up to the time where the
well reaches pseudo-steady state or steady-state conditions. Unlike production situations, steady-state behavior
is a real possibility with injection wells.

Infinite-acting flow regime: At early times wells do not sense reservoir boundaries. Flow during this period is
transient.

Pseudo-steady-state flow regime: Infinite-acting flow ends. After a transition period, a well in a closed
system feels the effects of the boundaries of the reservoir. After this transition period, pseudo-steady-state flow
develops. The pressure changes linearly with time. The change in pressure, dp/dt, is constant at all points in the
reservoir. This flow regime occurs in all closed reservoirs.

Steady-state flow regime. Steady-sate flow occurs when the pressure at every point in a system does not
vary with time. This situation can only occur in reservoirs that are completely recharged by a strong aquifer or
where injection and production are balanced (i.e., in certain flooding and injection programs). If there are
situations where steady-state flow exists, the relationships are only slightly different from those for pseudo-
steady-state flow (the 0.75 in Equation 1 is replaced by 0.50).

It may be necessary to estimate the injection rates before pseudo-steady state conditions develop. It is possible
to estimate the time to pseudo-steady-state production and the flow conditions during transient behavior. For an
injected fluid identical to the reservoir fluid, the time needed to reach pseudo-steady-state can be estimated as:

(2)

where:
tDA-pss = function of the reservoir shape (tDA-pss = 0.1 for a well at the center of a
bounded circular reservoir; (see Earlougher, 1977, Table C-1)
A = 2 2
area, for a circular reservoir, A = pre (feet )
tpss = time to pseudo-steady-state flow (hours)
k = permeability (md)
f = porosity (fraction)
ct = -1
total system compressibility (psi )

Prior to pseudo-steady-state, for infinite-acting radial flow, the injection can be estimated from the following
equation (ignoring turbulence):

(3)
where:
t = time (hours)

Using the above equations for a circular reservoir, the approximate time for pseudo-steady state flow to develop
is given by:

(4)

Before pseudo-steady-state (Equation 2 is an estimate of the time needed to reach a pseudo-steady-state


regime), estimate flow conditions using Equation 3. After pseudo-steady-state conditions have developed,
estimate flow using Equation 1 or its steady-state equivalent.

RECAP (ignoring turbulence)


Flow Description Liquid Gas
Regime
Infinite Occurring at
Acting early times
Radial and
Flow boundaries
IAR are not
"felt."
tpss Time for
pseudo-
steady-state
flow to
develop.

Pseudo- IAR flow has


Steady- ended and
State the change
in pressure
with time,
dp/dt, is
constant
everywhere
in the
reservoir.
Steady- This regime
State only occurs
when there
is pressure
support at
the outer
boundary

.
Injection Performance Relationships Single Phase
As you can see in the previous equations, pressure change in the reservoir and injection are related by Darcys
Law. The injection into a reservoir, as a function of the reservoir parameters and the pressure differential can be
characterized over a range of pressure differences (difference between the bottomhole pressure and the
reservoir pressure). In its simplest form, this is a straight line (assuming single phase flow and using Darcys Law
for steady-state conditions), until fracturing occurs, as shown in Figure 1.

Figure 1. The performance relationship for a characteristic well for single phase, laminar flow of a
slightly compressible liquid. This example is for production. Injection is analogous.

By assuming pseudo-steady-state or steady-state radial flow, the performance curve for an injector above the
bubble point (the pressure is high enough that gas is in solution) can be developed from (same as Equation 1):

(5)

Figure 1 represents a well with water injection. Similar relationships are available for gas.
The curve shown (Figures 1) does not reflect restrictions imposed by the completion itself and any tubulars or
damage/stimulation (skin). These curves provide a starting point for determining the flow and pressure
conditions that you will encounter during your injection operations.

The Effect of Skin on Performance


You can see from the basic equations that performance relationships are affected by skin damage or flow
restriction or flow improvement. According to Earlougher, There are several ways to quantify damage or
improvement in operating (producing or injecting) wells. A favored method represents the wellbore condition by
a steady-state pressure drop at the wellface in addition to the normal pressure drop in the reservoir. The
additional pressure drop, called the skin effect, occurs in an infinitesimally thin skin zone [for mathematical
purposes even though damage may extend deep into the formation]. In the flow equation , the degree of
damage (or improvement) is expressed in terms of a skin factor" (s) which is positive for damage and negative
for improvement. It can vary from about 5 for a hydraulically fractured well to + for a well that is too badly
damaged to produce (Earlougher, 1977).

The skin factor itself is dimensionless, but you can have an indication of the supplementary pressure drop caused
by skin (relative to an undamaged well) for radial, non-turbulent liquid flow using:

(6)

where:
Dps = pressure drop due to skin (psi)

It is important to consider the skin factor that the completion will develop, how much skin can be alleviated by
manipulating the pressure during completion, and in the longer term, deciding when a well will need to be
worked over or abandoned. Alternatively to Equation 6, you can calculate the skin by specifying a finite depth of
damage in the formation using:

(7)

where:
ks = permeability of the damaged zone (md)
rs = radial extent of the damaged zone (feet)

Unfortunately, the depth of damage and the damaged zone permeability are hard to infer. You can also assess
the influence of skin by considering skin in terms of an apparent (or effective) wellbore radius:
(8)

where:
rwa = apparent wellbore radius. Positive skin implies that the well has been damaged and the apparent wellbore
radius is reduced, and vice versa for a stimulated well (feet).

This relationship implies a reduced apparent wellbore radius (and consequently less production) for positive
skin. Figure 2 shows skin schematically.
Figure 2. Schematic representation of the variation of skin as a pressure drop through the reservoir,
for a particular flow rate. With positive skin more pressure is required to achieve a
particular rate. Note that this shows the real variation of damage/improvement away
from the well. Some of the mathematical treatments represent this pressure change as
occurring in an infinitely thin zone at the wellbore. This figure is for production. Similar
figures are possible for injection.

Skin can develop in a number of different ways, some depending on the damage done to the formation during
drilling and/or completion, some due to the completion itself, and some even due strictly to the inclination of the
well. The total skin is the combination of mechanical and pseudo-skins (total skin is determined from well testing
analysis).

Mechanical skin is associated with formation damage due to drilling fluids, cementing, etc. Mechanical skin is
mathematically defined as an infinitely thin zone that creates a steady-state pressure drop at the formation face.
A primary goal is to minimize turbulence, multiphase effects, perforation losses and restrictions due to
completion hardware.

The total skin can be considered as a summation of skins from various mechanisms, some of which can be
impacted by underbalanced operations. For example:

(9)

where:
s = total skin (all skins are dimensionless)
sM = mechanical skin (due to damage)
sPP = partial penetration skin (only part of a zone is completed)
sQ = rate-dependent skin (due to turbulence)
sPERF = skin associated with perforating or completing

Mechanical Skin (Drilling/Completion Damage)


These skins develop during drilling, workover or injection/production and are associated with the formation near
the wellbore. Undamaged, naturally fractured reservoirs can have a slightly negative skin (anything beyond 3
is probably induced). These reservoirs are extremely susceptible to drilling and completion damage.

Consider the situation shown in Figure 4, where concentric radial damage has been created by drilling and/or
completing overbalanced. How do you characterize the skin?

Lets suppose that we are dealing with permeable flow alone and that there is no damage or improvement
around the well.
The mechanical skin for such a case is 0.

The only pressure drop is due to flow through the porous medium. Denote this as .
If there is skin present, the bottomhole injection pressure, for the same rate as the undamaged case, will be
lower or higher (stimulated or damaged, respectively). Lets denote this as pwfs.
When there is skin, the pressure drop from the undamaged part of the formation (through the "damaged"
zone) into the wellbore would be .

You can see that the difference between Dpk and Dpks indicates the pressure drop due to mechanical skin,
giving Dps = Dpks - Dpk
If you use Equation (6) you can estimate the skin as (laminar):

If you know, or can estimate, rs, you can estimate the permeability in the damaged zone by using Dps in a
steady-state version of Darcys law:

(10)

If you know, or can estimate, rs , Equation 7 can be used.

-s
Finally, you can consider an equivalent wellbore radius, using Equation 8, rwa = rw

Partial Penetration
Partial penetration or partial completion skin is associated with drilling through only part, or completing only
part, of a formation. Though it may be most important during early time, it can be an issue through the lifetime
of a reservoir. This skin depends on the deviation of the well through the target zone, the amount of formation
penetrated during drilling and/or the amount of formation open to the wellbore.

Rate-Dependent Skin
Rate-dependent skin is associated with turbulent flow. Turbulent flow can be important in high-rate wells
particularly for gas but also oil in some high rate situations. If this is the case, the flow equation is modified to
represent an additional pressure drop. To account for turbulence in the previously developed relationships (see
Equation 1):

(11)

For single phase liquid flow:

where:
ro = oil density at reservoir temperature and average pressure .

For additional reading, refer to:


Jones, L.G., Blount, E.M., and Glaze, O. H.: Use of Short Term Multiple Rate Tests to Predict
st
Performance of Wells Having Turbulence, SPE 6133, SPE 51 Annual Fall Meeting, New
Orleans, LA (1976).

Figure 3. Schematic representation of mechanical skin in a generic reservoir

Perforation/Completion Friction
Perforation friction can also impair injection. Perforation friction - pressure drop due to flowing through the
perforated, or other, completion - can definitely be reduced by appropriate underbalanced completion, in addition
to an intelligent perforating program. According to Suman, Perforating is always a cause of additional damage in
4
formation rocks. Whether it is perforated overbalanced or underbalanced, perforating compacts the formation
around the perforations. This compacted zone probably has an average thickness of 0.5 inches and the
permeability reduction averages 80%.

While compaction damage during perforating is anticipated (regardless of the balance conditions), underbalanced
perforating offers significant advantages for removal of gun debris and damaged material and may reduce the
extent of the damaged zone. You can estimate pressure drop associated with a perforated completion and infer
the corresponding skin factor using the methods described in Completion Hydraulics.

Typically, pressure drop and skin estimates are based on the premise that all guns fired properly and that all
perforations are open. You can infer the damage due to alone by measuring the total skin with some form of well
testing, isolating the skin due to turbulence and then subtracting the skin due to perforating, partial penetration
and wellbore geometry. Perforation friction is often not adequately considered. Though it is often rather small,
there can be important exceptions, and you should include it in your completions' design.

For now, consider the implications of the relationship shown in Figure 4. This is the most basic perforation design
plot, highlighting the role of spacing. The trends are as you would expect greater perforation density generally
leads to higher rates. Design methodologies are available for spf, phasing, charge size, gun characteristics and
balance. One of the more difficult characteristics of a perforating program is to estimate how much the near-
perforation permeability is damaged and to what extent. In a well consolidated formation, the damaged depth
can be 0.5 inches and permeability may be reduced to 10 to 25% of the virgin formation permeability (McLeod,
1983).

Performance relationships are impacted by skin. It is important to minimize formation damage during all phases
of drilling, completion, and injection. For example, Figure 1 is based on no skin, and Figure 5 shows the
influence of skin on performance.

Figure 4. Schematic relationship showing the variation of total inflow with perforation density. This
curve is schematic. The particular curve shape depends on the perforation and formation
specifics and the degree of underbalance during shooting.

Figure 5. Example IPR curves for an oil well (ignoring dissolved gas for now) with different values
of total skin. Even modest skins can seriously impact productivity and overall economics
(i.e., the bottomhole flowing pressure must be lower to achieve the same rate as the skin
increases). This example is for production.

Gas Flow and Turbulence


Flow mechanics in the reservoir, in the completions, and in the production string are greatly complicated when
turbulent flow occurs. For a dry gas or for a situation with a GOR > 30,000, recall that pseudo-steady-state gas
flow can be expressed as:

(12)
where:
qg = gas flow rate (Mscf/D),
k = permeability (md),
h = thickness (feet),
= average temperature (R, R = F + 459.67), and,
= average compressibility factor (dimensionless) (Average viscosity should also be
considered.)

This equation can be written as:


(13)

It is helpful to separate pressure drop due to turbulence from skin arising from other sources so that you can
assess the specifics of the completion. The quadratic term on the right hand side of Equation 13 accounts for
non-Darcy, turbulent effects. This relationship can be rearranged by dividing both sides by the flow rate:

(14)

This equation is sometimes a more convenient form for visualization because it can be expressed as a straight
line (as shown in Figure 8), and the term on the left-hand side of this equation can be considered as a reciprocal
injectivity or productivity index.
Figure 6. Variation of injectivity/productivity index for gas as a function of the gas flow rate.

This type of plot is useful for identifying the presence of turbulence. If this plot has a positive slope (as shown in
Figure 6), turbulence is indicated. Figure 7 shows how this type of plot can be useful for discriminating between
turbulence and other types of skin. In Figure 7, relationships are shown for eight generic well situations in a
hypothetical field different wells are indicated by the numbers on the plot. Wells 1 and 5 show no turbulent
effects because b = 0 (slope is zero). Wells 2 through 4 and 6 through 8 show increasing levels of turbulence.
Why are the two sets of curves different? There is greater skin (not associated with turbulence) for Wells 5
through 8.

Figure 7. Variation of injectivity/productivity index for gas as a function of the gas flow rate, with
varying degrees of turbulence and skin due to damage.

Evaluating the skin due to turbulence helps you to determine whether the pressure drop is intrinsic to the
formation and/or the completion and cannot be overcome without stimulation. You can use simple diagnostic
plots to separate skin caused by turbulent flow from skin caused by damage and, to a certain extent, skin caused
by the completion itself. Similarly, you can use simple methods to design the most effective completion
(beforehand) and to assess its effectiveness or the need for remedial measures (after the fact).
You can prepare the same types of plots for oil to separate turbulence from other skins (turbulence is less
common when oil is produced, but it can occur). Recall that the slope b in Figure 7 is indicative of the nature of
the skin in the well. If b is high, it indicates high turbulence. A large value of the intercept a implies that the skin
is high (not associated with turbulence) or that permeability is low.

If the well has been fractured, you must consider other factors.

For additional reading on this subject refer to: Beggs, H.D.: Production Optimization Using
TM
Nodal Analysis, OGCI Publications, Oil & Gas Consultants International Inc., Tulsa, OK
(1991).

Horizontal Wells

The relationships presented so far are for straight hole. Various relationships have been extended to approximate
flow in horizontal wells. Be careful. Analytical relationships may be inadequate (A. Settari, personal
communication). Some of the analysis techniques are summarized below.

Using numerical simulation, Bendakhlia and Aziz (1989) suggested that:


(15)

where:
V = variable parameter in a modified Vogels equation
n = exponent in Fetkovichs equation

To use this, recognize that there are three unknowns (qmax, V, and n). At least three stabilized flow
measurements are required to build the IPR.

For pseudo-steady-state flow, Joshi (1991) presented an approximate formulation:


(16)

where:
JH = Productivity/Injectivity Index (PI) for a horizontal well section (BOPD/psi)
h = true vertical thickness of a zone (feet)
reH = drainage radius (feet)
kh = horizontal permeability (md)
kv = vertical permeability (md)
L = horizontal length (feet)
rw = wellbore radius (feet)
b = (kh/kv)
0.5
Thomas et al. (1996) showed calculations of near-wellbore skin and the non-Darcy flow coefficient (turbulence)
for horizontal wells for various drilling and completion scenarios. They provided analytical and numerical
solutions for near-wellbore skin and non-Darcy flow for horizontals drilled underbalanced or overbalanced with
various completion options. In particular, they explored the effects of drilling overbalanced versus
underbalanced and completing openhole with or without a slotted liner or completing cased hole. This reference
is recommended.

According to Thomas et al., Traditional formulations of wellbore skin assume radial flow into a vertical wellbore,
and must be transformed to apply to horizontal wells. In these formulations, adopt the convention that the
horizontal wellbore is oriented with the x-axis and its length is L. k is the geometric mean of the two permeability
components perpendicular to the direction of the well. Implementation of these terms [skin terms] into different
wellbore models results in different multiplying factors on the skin and non-Darcy flow terms depending upon
how a particular well model was derived. This approach is different than the treatment for vertical wells (Thomas
et al., 1992) where the effect of partial penetration on near-wellbore skin is included with the s and D terms.

Available formulations, by various different authors, are shown below. Thomas argued that there were basically
no significant predictive differences between any of the models.

Mutalik, et al., (1988) developed a relationship by using the solution for a fully penetrating infinite conductivity
fracture. For pseudo-steady state flow into a horizontal well:

(17)

where:
sf = skin for a fully penetrating vertical fracture (dimensionless)
c = horizontal shape factor in this model (dimensionless)
r e = effective drainage radius (feet)
b = a turbulence factor
s = mechanical skin (dimensionless)
kx = horizontal permeability in the direction of the well (md)
ky = horizontal permeability normal to the direction of the well (md)
kz = is the vertical reservoir permeability (md)
L = is the length of the completed interval (feet)
sCa,h = is a skin associated with the drainage areas shape (dimensionless)

Babu and Odeh (1989) developed an equation from the classical vertical solution turned on its side, accounting
for the resulting geometry. Thomas et al. (1996) added mechanical skin to this relationship to give the following
equation:

(18)

where:
sR = partial penetration skin in the Babu and Odeh model (dimensionless)
xe = half-length of the drainage area in the x-direction (feet)
Al = horizontal well drainage area (ft )
2
CH = horizontal shape factor in the Babu and Odeh model

Economides et al. (1994) used a semi-analytical technique in which an instantaneous point source analytical
solution was integrated numerically in time and space to give constant flux solutions for a horizontal well located
anywhere in the drainage volume of a uniformly heterogeneous reservoir with three-dimensional permeability
anisotropy. They used a horizontal shape factor from Besson (1990). These solutions are used to calculate
constant pressure solutions for pseudo-steady-state:

(19)

where:
k = absolute permeability (md)
xe = drainage length in the x-direction (this is the well direction)(feet)
sx = vertical skin effect (Kuchuk, 1995) (dimensionless)
se = term to account for vertical eccentricity (dimensionless)
sew = distance of the well from the center of the reservoir (feet)
Ch = a horizontal shape factor in the Economides et al. model (dimensionless)

There are certainly other analytical formulations. For additional information, the reader can review Goode and
Kuchuk, or Kuppe and Settari, 1996.

Analytical and even numerical prediction of the behavior in horizontal wells can be quite difficult. As Teichrob et
al. (1998) state: Without some knowledge regarding inflow distribution and magnitude, accurately predicting
bottomhole pressure becomes problematic at best and impossible at worst. At best, design considerations must
focus on dead or non-productive holes, then attempt to bracket several flowing conditions recognizing differing
points of inflow and magnitude. A reasonable integrated approach can use information acquired during
underbalanced drilling. According to Butler et al. (1996). For a number of years we have postulated that sweet
spots within a lateral section of a reservoir could be identified (in real time while drilling) through superposition
of elemental inflow performance relationships. Using 50 meters as [an] elemental wellbore length, the concept is
as follows: element #1 is drilled and inflow is measured and correlated to a given pressure drop. Injection ratios
are changed, while still honoring hole cleaning and bottomhole pressure constraints. New inflow rates are
measured and correlated to a different pressure drop. An inflow relationship is therefore developed for element
#1, which describes inflow over a range of differing pressures. A new element is drilled. Because we have a
relationship that describes inflow over a range of pressure drawdown for element #1, we are able to reconcile
what the contribution of that previous element should be based on the new pressure drops across the first
element. Overall, flow to surface can be measured and will be the sum of predicted inflow from element #1 plus
inflow from the newly drilled element.

Multilaterals
Following horizontal wells, the next most recent trend in reservoir exploitation has been to drill multilaterals.
Some inflow performance procedures have been formulated for multilaterals and branched wells (for example,
Larsen, 1996). These can be useful in determining if underbalanced completion is merited.

Larsen used pseudo-radial skin factors for systems of connected wells, intervals or fractures with a common
bottomhole pressure or potential. The approach assumes a certain distance between the midpoints of the well
elements (e.g. branches, drainholes, fractures, the well, or multiple wells) and then assumes that the pressure
development only depends on the distance. Larsen stated, Even relatively close to a horizontal well, it is, for
instance, possible to estimate the pressure development with negligible error by using a fully penetrating vertical
line well. Moreover, since the line-source solution can be approximated with a simple logarithmic expression at
sufficiently later times, it is possible to combine the effect of, for instance, several wells or branches by
elementary methods.

Larsen argued that you can apply the basic pseudo-steady state equations (e.g. Equation 2 for a circular,
bounded drainage area) to any well configuration with some minimal distance between inner and outer
boundaries of the system. For cases with significant boundary effects occurring before pseudo-steady-state flow
(e.g., for a horizontal well approaching the length of the reservoir in the direction of the well), you should use a
direct solution rather than analytical approximations. Regardless, for situations where the equivalent wellbore
radius remains within the drainage area, indirect methods are acceptable.

By considering symmetrical characteristics in the well system, Larsen implemented the basic analytical
relationships for wells with long horizontal branches in bounded reservoirs. For regular patterns of branches or
wells treated as a single drainage system, Larsen suggested subdividing and adding the partial IIs. For nw wells:

(20)

According to Larsen, For wells with multiple branches, the pressure transient behavior can be obtained by
spatial superposition of solutions for individual branches. The late time difference from the solution of a fully
penetrating vertical well can then be used to determine the pseudo-radial system skin factor of the given well
configuration. This analysis is complicated for layered or heterogeneous reservoirs and at this time it should
consider only single layered systems with horizontal isotropy.

Other analytical formulations for evaluating transient behavior in multiple laterals are available in the literature
(for example, Ozkan et al., 1997).

Obviously, these situations are so complex that simulation is recommended. However, for back-of-the-envelope
evaluations, methodologies like this can be used for determining productivity, for building composite
performance curves, and for evaluating the contribution of mechanical skin (which may be alleviated by
underbalanced operations, if economically appropriate).

Transient Performance Curves


Lets take a step back. We have briefly summarized some of the issues for relationships between pseudo-steady-
state or steady-state flow. The complexities for pseudo-steady-state flow in horizontal wells have been indicated.
You can see that developing performance curves for multilaterals during pseudo-steady-state flow is even more
complicated.

Most of the relationships shown to this point have been for pseudo-steady-state or steady-state flow. When a
reservoir is produced without voidage replacement, the average reservoir pressure is reduced (i.e., transient
pressure behavior) and the inflows performance changes. The performance can also change with time even if the
average reservoir pressure does not change significantly - the flow behavior is transient. During underbalanced
drilling and completion, flow usually is transient, but you can develop IPR curves for any time in the productive
life of a reservoir, as shown in Figure 8.

Wells initially start out under infinite acting radial flow (unless the perforations or completed zone are very
restrictive spherical flow). As the radius of influence approaches the boundaries, the flow regime changes until
pseudo-steady-state is reached. Further behavior depends on the type of boundary (e.g., no flow, constant
pressure, mixed, etc.). Regardless, most drilling and completion scenarios can probably be reasonably
represented as infinite-acting radial flow. This flow condition in itself is transient.

Equation 3 showed relationships for how flow and/or pressure change during this transient period. These are:

(21)

Figure 8. Typical transient performance curves for a reduction in average reservoir pressure. You
can see that pressure-rate relationships change over the producing time. There are
implications. For example, behavior during a workover intervention may be different than
during the initial completion. This example is for production.

Regardless of which technique you choose, you can develop IPR curves for various values of drawdown,
depletion, and time. Remember that m, B, ct, and change with reservoir pressure and drawdown.
Summary
Most injection operations, but not all, will be in pseudo-steady-state or steady-state regime. This is all
premised on the injection fluid properties being the same as the reservoir fluid properties.

If your reservoir is high permeability the period for infinite acting radial flow (IAR) may be short.

As soon as the pressure disturbance caused by flow from the well reaches the outer boundary, a transition
period starts from infinite acting radial flow, going to a pseudo-steady-state period, or to steady-state.

The pseudo-steady-state period occurs if a no-flow boundary farthest from the well is reached by the
pressure disturbance and the entire area of influence is affected by injection. Rate and wellbore pressure tend
to stabilize.

When does pseudo-steady-state behavior start? As was indicated in Equation 2:

(22)

Can I develop performance curves for infinite-acting conditions as a series of rate-pressure plots just using
Equation 21 and inputting different values of pwf? You can only do this to provide rough indications of
behavior because the transient effects are not properly accounted for when the rate changes.

Recommended Reference: Golan, M., and Whitson, C.H.: Well Performance, Second Edition,
Prentice Hall, Englewood Cliffs, NJ (1991).

What Happens if the Reservoir is Layered?


In many cases, the reservoir encountered may not be homogeneous. Consider the simple situation shown in
Figure 9. The permeability can be approximated as:

(23)

Figure 9. Schematic of a layered reservoir.


where:
= average permeability (md)

h = gross thickness (feet)


i = subscript denoting an individual layer

The situation is dramatically more complicated if there is variable water or gas cut, different rates of depletion in
different zones, and crossflow. This is not to say that there are not some possible approximations. Consider
commingled injection (no vertical, in-reservoir crossflow) of two zones. If the reservoir pressure in Zone 2 is less
than that in Zone 1, net production will not occur until the producing pressure, pwf is less than a critical value,
p*wf. Above this pressure, fluid just goes from one zone to the other. This critical pressure is:

(24)

For this basic commingled case, you are able to construct the performance curve for each zone and add the
predicted rates at various pressures less than p*wf. Somehow, however, you have to know the properties of
each zone (core analysis, logging, production logging, PTA) from this well or from offsets.

The previous sections provided a brief introduction to some of the reservoir engineering considerations that are
essential to effectively and safely design any injection. Note that this only covers matrix injection without a
front. But the performance of the reservoir is only part of the whole story. The other half of the story is that flow
to the surface can be restricted by the specific completion, the string in the hole, and the surface facilities. This
relates to the hydraulic behavior of the completion itself and the tubing. Completion hydraulics are described
first and then tubing-related hydraulics are briefly covered.

Completion Hydraulics

After evaluating the pressure profile in the reservoir, the next step in estimating potential well performance is to
estimate the pressure drop through the specific completion. The scenarios can range from barefoot to gravel
pack completions. You can assess barefoot completions using the methods already described. This section covers
both gravel pack and perforated completions. Gravel pack completions are described for completeness and for
the possibility that they can be components in underbalanced interventions. You can also use these examples to
infer skins associated with slotted liners, screens, etc.

Cased-Hole Gravel Pack Completion


The brief background in this section will help you in assessing the comparative merits of this type of completion
and in estimating skin. You can also modify the basic concepts for situations where other sand exclusion
hardware is run (e.g., slotted liners, expandable sand screens, etc.). For example, Kaiser et al, 2000 provide
data on slotted liners and Asadi and Penny, 2000 describe pressure drop through screens.

In its simplest terms, the pressure drop that occurs across a gravel pack completion is related to the gravel
permeability (and damage), the perforation density, length and damage. You can estimate the pressure drop
(and ultimately convert it to a skin, if appropriate) using the following equations:

For Oil or Water:

(25)
where:

pwfs = sandface flowing pressure determined from the IPR (psia)


pwf = bottomhole flowing pressure (psia)
q = flowrate (BLPD)
m = viscosity (cP)
B = formation volume factor (RB/STB)
Lperf = distance from the outer screen diameter to the tip of the perforation (feet)
kgravel = permeability of the gravel (md) (see Table 1)
Aperf = available area for access to the formation (total of the nominal perforation
2
cross-sectional areas through the perforated interval) (feet )
bgravel = a turbulence term for the gravel
r = 3
liquid density (lbm/ft )
hperf = net perforated interval (feet)
nperf = shot density (spf)
dperf = perforation diameter (inches) (see Table 2)
dscreen = screen diameter (inches) (see Table 3)

Table 1. Typical Gravel Permeability


Gravel Size Average Porosity Average Permeability
(U.S. Mesh) (%) (darcies)
40-60 39.8 69
20-40 40.9 171
10-20 40.5 652
8-12 41.5 1969
6-10 42.0 2703
Note: These are representative numbers. Confirm all values with the
appropriate service companies. This is actually quite important. Three different
sources gave three different permeability ranges. The values shown here can be
used for default calculations. They are from Dowell Schlumbergers Well Analysis
Manual, circa 1986.
Table 2. Typical Perforation Information
Gun Size Casing Size Average Perforation Penetration
(inches) (inches) Diameter (inches)
(inches)
Average Longest
Through Tubing Retrievable
0.21 3.0 3.3
1 4 casing
0.24 4.7 5.4
1 5 casing
0.24 4.8 5.5
1 4 to 5 casing
2 0.32 6.5 8.1
4 to 5 casing
0.33 7.2 8.1
2 2 tubing to 4 casing
0.36 10.3 10.3
2 4 casing
Through Tubing Expendable
0.19 3.1 3.1
1 4 casing
0.30 3.9 3.9
1 2 tubing??
0.34 6.0 8.1
1 2 tubing to 5 casing
0.42 8.2 8.6
2 5 to 7 casing
0.39 7.7 8..6
2 2 tubing
Retrievable Casing Guns
0.38 10.5 10.5
2 4 casing
0.37 10.6 10.6
2 4 casing

0.42 8.6 11.1


3 4 casing
0.36 9.1 10.8
3 4 casing
0.39 8.9 12.8
3 4 and 5 casing
4 0.51 10.6 13.5
5 to 9 casing
5 0.73 12.3 13.6
6 to 9 casing
NOTE: These are representative numbers only. Confirm all values with the appropriate service
companies. In fact, many of the perforating companies now have calculation programs for
approximating diameter and depth of penetration. This table should be used as a fallback
resource only.

Table 3. Common Screen Diameters for Inside Casing Gravel Packs


Casing Maximum Screen Commonly Used Screen
Diameter Diameter*
OD Weight ID Pipe OD Wire OD Pipe OD Wire OD
(inches) (lbf/ft) (inches) (inches) (inches) (inches) (inches)
4 9.5 3.548 1 1.815 1 1.815
11.6 4.000 2.160 2.160
4 1 1
5 18.0 4.276 2.400 2.400
1 1
17.0 4.892 2.875 2.875
5 2 2
24.0 5.921 4.000 3.375
6 3 2
7 29.0 6.184 4.000 3.375
3 2
33.7 6.765 4 4.500 3.375
7 2
36.0 7.825 5 5.500 3.375
8 2
47.0 8.681 6.000 3.375
9 5 2
NOTE: These are representative numbers. Confirm all values with the appropriate service
companies. This table should be used as a fallback resource only.
*Different screen diameters may be required for specific circumstances such as larger diameter
production tubing, multiple gravel packs, etc.

For Gas
It is commonly assumed that most of the pressure drop occurs in the packed perforations (at least as a starting
point). As will be discussed later, this is likely not completely true. Nevertheless, assuming that this is the case,
linear flow in the perforation tunnels can be represented as:
(26)

where:
pwfs = sandface flowing pressure determined from the IPR or deliverability equations
(psia)
pwf = bottomhole flowing pressure (psia)
q = volumetric flowrate (MscfD)
m = gas viscosity at the mean pressure (cP)
=

mean pressure, = (psia)


= real gas deviation factor (dimensionless) at T,
T = temperature (R = F + 460)
Lperf = distance from the outer screen diameter (feet) to the tip of the perforation (feet)
dscreen = screen diameter (inches) (see Table 3)
kgravel = permeability of the gravel (md) (see Table 1)
Aperf = available area for access to the formation (the total of the nominal perforation
2
cross-sectional areas through the perforated interval) (feet )
bgravel = a turbulence term for the gravel
hperf = net perforated interval (feet)
nperf = shot density (spf)
dperf = perforation diameter (inches) (see Table 2)
gg = gas gravity (dimensionless, air = 1.0)

Example: Calculating IPR for a Cased-Hole Gravel Pack Completion


The frictional pressure drop through the completion (which can be converted to a skin) can be determined as a
function of the rate. Consider an example for an oil reservoir injecting into a wellbore and the injected fluid has
the identical properties to the fluid in the reservoir. The relevant parameters are:

Average reservoir pressure = 2,000 psia


Water viscosity at reservoir temperature 1 cP
Water formation volume factor = 1 RB/STB

Reservoir permeability = 100 md


This well has a drilled diameter of 9 inches
re ~ 1140 feet

Calculate Performance Relationship for this Example as Follows:


1 Suppose that size selection evaluation indicated that 10/20 gravel was appropriate. For 10/20 gravel
(from Table 1):

kgravel = 652 darcies, .

2 Suppose that the well is vertical, the perforated interval is 30 feet, the shot density is 24 spf, and a
5-inch casing gun will be used in 6-5/8 inch, 20 lb/ft casing. For this gun, the following specifications
apply (see Table 2 for approximate penetration features):
dperf 0.73 inches
Lperf {12.3 inches = 1.025 feet} [perforation penetration]
+{6.049 inches = 0.504 feet}/2 [radius of 6 5/8-inch, 20 lb/ft casing] - {3.375
inches = 0.281 feet}/2 [outer screen radius]
= 1.14 feet
3 Calculate the available area for access to the formation as:
-3 2 2 2
A = 5.454 x 10 x 30 feet x 24 spf x 0.73 in = 2.093 feet
4 Determine the linear coefficient in the flow equation as:

5 Calculate the quadratic coefficient in the flow equation as:

6 The pressure drop through the completion is:


-7 2
pwf - pwfs = 0.00074q + 1.37 x 10 q

7 Construct a performance curve for the reservoir itself using the equations presented previously. These
data are shown in the first two columns in Table 3.

8 Determine the pressure drop due to the completion (pwfs - pwf in Table 4). Substitute the rate from the
second column in Table 4 into the equation in Item 6 above to calculate pwf - pwfs

The pressure drops due to the completion itself can be plotted to indicate the contribution from the completion
more specifically, as shown in Figure 10.

Figure 10. Pressure drop due to the cased-hole, gravel pack completion in the example case. The
baseline reservoir performance assumed no mechanical skin (no plugging), but
formation damage skin can be considered. Just use a skin term when the formation
performance is calculated.

Cased, Cemented and Perforated Completion


The previous gravel pack completion example included some obvious simplifications about the effectiveness of
the perforations. It assumed that the perforations themselves did not damage the formation in any way. In
reality, the skin associated with perforating can be surprisingly important. Perforation pressure drop (from which
skin can be calculated) can be inferred. Some approximations for a cased-cemented-perforated completion are
shown below.

For Oil or Water


The procedure is not unlike that taken for the cased hole gravel pack, with the exception of the presumed flow
regime around the perforation tunnel. Assuming that there is no interaction between perforations, a first-order
assumption is that there is radial flow from each perforation. Assuming this:
(27)

where:
Dp = pressure differential across the completion (through the perforation and the
surrounding crushed/damaged zone) (psi)
q = flow rate per perforation (BOPD/perforation)
qtotal = total flow rate (BOPD)
b = turbulence factor
rperf = perforation radius (feet)
dperf = perforation diameter (feet)
nperf = perforation density (spf)
hperf = net perforated length (feet). Partial penetration of a deviated well will create
additional skin
m = viscosity (cP)
Bo = Formation Volume Factor (RB/STB)
rc = radial extent of the crushed zone from the centerline of the perforation tunnel
(feet). The example shown is for a situation where it is assumed that the crushed
zone around the tunnel is 0.5 inches thick. You can use other values, if available.
This is a common default assumption. As can be seen in the equation, the crushed
zone thickness is converted to feet by dividing by 12.
kcf = permeability of the perforation damaged zone (md). This is determined by
multiplying the native reservoir permeability (k) by a damage factor.
ro = 3
oil density (lbm/ft )

The permeability of the compacted zone is commonly assumed to be a fraction of the nominal reservoir
permeability. A rule-of-thumb is:
kc = 0.1k Overbalanced (28)
kc = 0.4k Underbalanced
This in itself shows a key reason for perforating underbalanced the degree of damage in the compacted zone
can be significantly less than when perforating is done balanced or overbalanced.

For Gas
The approximations shown above are a generic and ideal approximation for liquid flow through perforations.
Similar concepts can be applied for gas. If the perforations are considered to not interfere with each other, a
first-order assumption is that there is radial flow from each perforation. Based on this assumption:
(29)

where:
q = flow rate per perforation (MscfD/perforation)
qtotal = total flow rate (MscfD)
nperf = perforation density (spf)
hperf = net perforated length (feet)
m = viscosity (cP)
b = turbulence factor
rperf = perforation radius (feet)
dperf = perforation diameter (feet)
rc = radial extent of the crushed zone from the centerline of the perforation tunnel
(feet). The example shown is for a situation where it is assumed that the crushed
zone around the tunnel is 0.5 inches thick. You can use other values, if available.
This is a common default assumption. As can be seen in the equation, the crushed
zone thickness is converted to feet by dividing by 12.
Lperf = perforation length (feet)
kc = permeability of the perforation damaged zone (md). This is determined by
multiplying the native reservoir permeability (k) by a damage factor. The same
default assumptions for the damage factor are made as for oil.
gg = gas gravity (dimensionless, air = 1)
= mean real gas deviation factor (dimensionless)

T = Temperature (R = F + 460)

Calculating Performance for a Cased and Perforated Oil Well


Consider the example (an injector) shown in Table 4. Assume that this was shot underbalanced.
Table 4. Basic Reservoir and Completions Data
Parameter Value Calculations
Average reservoir pressure 2000 psi
Water viscosity 1 cP
Water formation volume 1 RB/STB
factor
Water Density 54.67lbm/ft
3

Well Orientation vertical


Perforation Density 12 spf or 4 spf
Perforated Interval 30 feet
Casing 6.625"
Gun 5" casing gun
(underbalanced)

Permeability 100 md
Drainage Radius 1140 feet
Bubble Point Pressure 1200 psi
Drilled Diameter 9 inches

1 From these data, determine the linear coefficient in the flow equation as:

2 Calculate the turbulence factor as:

3 Calculate the quadratic coefficient in the flow equation as:

4 Determine the pressure drop through the completion using:


-3 2
pwf - pwfs = 2.97q + 6.31 x 10 q

5 Dont forget that the performance is calculated with the total flow rate and that the relationship for the
perforations is based on flow per perforation. Pressure drops for 12 and 4 spf are shown.

The pressure drops due to the completion itself can be plotted to indicate the contribution from the completion
more specifically, as shown in Figure 11.
Figure 11. The pressure drops due to the completion in the example case. The baseline reservoir
performance assumed no skin, but formation damage skin can be considered.

Other Completions

Methods were shown for estimating the pressure drop for two generic completion types - a cased and perforated
completion and a gravel pack completion. The gravel pack calculations shown were for an ideal pack without any
associated damage to the perforations. For the gravel pack, the basis for the calculation was to assume that
there was linear flow along the perforation tunnel and through the gravel. This analysis did not incorporate any
damage to the pack or to the perforations. To accommodate damage to the gravel pack itself (either due to
placement or fines, etc.) the gravel pack permeability can be reduced. To account for perforation damage, split
the problem into two parts first, calculate the pressure drop for linear flow through the gravel in the pack and
in the perforation tunnel. Then, calculate the pressure drop through a compacted zone around a perforation as
shown in the second example. Then combine the pressure drops or skins.

These conceptual methods can provide guidelines for determining pressure drops through other types of
completions. Consider an openhole gravel pack. This situation could be considered as a two-part reservoir with
radial flow (one permeability zone is the gravel and the second is the formation itself). This problem can be
solved by using some of the formulas presented earlier. One approximation is to use:

For example, let k be the formation permeability with a value of 1000 md. Let ks be the gravel pack permeability
assume that the pack has been placed perfectly and the gravel has a permeability of 650 darcies. At this
point, be a little careful - use rw as the outer diameter of the screens. Use rs as the drilled diameter of the
wellbore. Lets assume a 9-inch hole, giving rs = 9/(2 x 12) = 0.375 feet. If a 3 3/8-inch diameter screen wire
wrapped screen is used (and pressure drop is ignored through the screen for demonstration purposes at least),
the artificial wellbore radius can be taken as 3.375/(2 x 12) = 0.141 feet. The skin would be:
Granted, this demonstration shows negligible skin. It becomes more significant if the formation permeability is
higher, if the gravel is damaged with placement or injection fluids, if the formation has been damaged by fluids,
solids and inadequately cleaned up, if plugging or other mechanisms are impairing the screens themselves and
so on.

Other types of completions provide specific pressure drops. Kaiser et al., 2000, provide an excellent
presentation of the methodologies that can be used for optimizing slotted liner selection.
The primary factors considered in [slotted liner] design are sand control, inflow resistance and cost. Inflow
performance is usually considered to be controlled by the open area exposed to the reservoir, and sand control
governed by slot opening size. These become competing considerations in reservoirs with fine sands because
slot density must be increased to maintain open area if slot size is reduced to control sand.

The pressure drop through this type of completion is not only a function of the slot, but also the flow
convergence in the sand that packs around the slots. In fact, the flow loss through an open slot is negligible
compared with that induced by the flow disturbance associated with the slot. You can visualize the difference
between uniform flow into a barefoot completion, with a slight convergence into the hole, with the much more
extreme convergence that concentrates flow into a slot at similar rates. These convergence effects can lead to
pressure drops that bear consideration during planning stages. You might think that fewer larger slots would
have less flow resistance than more smaller slots for the same open area. This presumption ignores what Kaiser
et al. indicate as the most important component of slot-induced loss this being the convergence of flow in the
sand that is packed around the slots. In fact, they indicate that this is why wire wrapped screen are so effective;
because the slots are very close together and this minimizes the extent of flow convergence and its associated
pressure loss. [Note that Kaiser was commenting on production.]

Kaiser and his colleagues outlined a combined empirical, analytical and numerical program where they optimized
liner characteristics for Marathons South Bolney field. Their analysis model combined the radial Darcy flow
relationships, a mechanical skin factor that is associated with flow convergence and flow along the completion
itself (see the section on Wellbore Hydraulics). They called the slot component of the skin factor the slot factor
and provided some illustrations of its relative contribution. The slot factors calculated from the flow
convergence analysis are dimensionless numbers that are used with the near-wellbore permeability to calculate
the pressure loss induced by the slots.

This paper is recommended reading. These authors also discussed the standard engineering considerations of
optimizing the completion to avoid slot plugging and the considerations that may be taken for evenly distributing
flow along a long slotted liner section in a horizontal or extended reach well.

Convergence through limited access in the completion hardware is the essential issue. In effect, this is
analogous to partial penetration. These effects can be compounded if blank sections are run within the liner
section. Blanks can be used for load control devices. The inflow performance is impacted by the reduction in the
net production interval. There is local flow convergence at the slots themselves and larger scale convergence to
the slotted intervals. Rough approximations of the partial coverage skin from this effect (for example, ignoring
axial flow) were provided as:

where:
B = the coverage ratio is B.
Manufacturers should be able to provide you with the skin that is associated with a specific completion. Of
course, plugging of the screens themselves compounds these skins. We all know that this is not a trivial issue.
A recent paper by Asadi and Penny, 2000 demonstrates some of these issues. One of the most important
aspects of this paper related to the methodologies and effectiveness of cleanup techniques on screens that were
damaged while being run this brings us back to some of the advantages of underbalanced operations where it
may be possible to mitigate the damage by running in certain less damaging fluids.

Standard sand control considerations are published elsewhere (for example, Penberthy and Shaughnessy,
1992.) The work of Asadi and Penny is valuable because large-scale empirical measurements were made on
commercial products (20/40 PMF porous fiber material equivalent to a 20/40 sand, 20/40 Stratapack, and
0.008-in gauge wire wrap Prepacked with 40/60 mesh resin coated sand). These components were exposed to a
formulated drill-in fluid containing drill solids. Various cleanup techniques were then attempted (10% HCl;
enzyme breaker followed by 10% HCl; acid activated breaker, followed by 10% HCl, with and without water
backflow. The efficiency of some of the evaluated hardware was low and cleaning was never 100% effective
sometimes much less. The products were vulnerable to plugging and cleanup under ideal laboratory conditions
was never ideal. Consider all methodologies for reducing the degree of plugging in the first place. Also
remember that the pressure drop even in clean screens will be elevated when turbulent flow exists. Asadi and
Penny, 2000, indicated pressure drops of between 0.05 and 0.1 psi at an injection rate of between 9 and 18
BLPD/ft of length. Turbulent flow will elevate these values even for clean screens.

The issue of turbulent pressure losses applies to all types of completions, including, as we have seen cased and
perforated situations. The reader may wish to review numerical work by Behie and Settari, 1993. Numerical
models such as these can also be used to evaluate the influence of collapsed and filled perforations another
potentially significant skin component even if shot underbalanced or maybe because they were, collapsed and
infilled perforations can add significantly to pressure drops that are experienced.

It all boils down to calculating skin. This section was intended to illustrate concepts. These days in most cases,
you would have software available. However, without understanding underlying concepts, using commercial or
internal software is inefficient at best, and possibly dangerous.

Using the concepts shown, you can estimate the degree of damage due to the completion and the skin due to the
completion itself. For an excellent summary of skin associated with perforating, refer to Chapter 6.5.3 in Bell et
al., 1995. For a complete discussion of sand control issues, refer to Penberthy and Shaughnessy, 1992. For
cogent discussions of the reservoir engineering aspects of various completions, refer to Beggs, 1991, and/or
Golan and Whitson, 1991).

There are methods for estimation of the pressure drop through the reservoir and the completion. To complete
the analysis of well performance you must next consider the effects of wellbore hydraulics.

Wellbore Hydraulics

There are methods for estimating pressure changes in the reservoir and across the completion. You will need to
complete the final component of the well performance analysis estimating the pressure drop in the wellbore
and at the surface. The pressure regime in the tubing or annulus is governed by hydrostatic, frictional, and
acceleration forces. This section shows you how to approximate these effects for single phase and multiphase
flow in the wellbore.

The key to estimating pressure drop is determining an accurate bottomhole pressure. Consider the simplest
situation single phase flow. You can estimate bottomhole pressure, (pBH or pwf) from surface pressure psurface
using the following equation:

pBH = psurface + pHydrostatic pFriction (30)


pHydrostatic is the hydrostatic head of the fluid in the tubing and/or annulus and pFriction is the loss in pressure due
to friction loss during flow.

The bottomhole pressure is important for:


Analyzing current performance
Predicting future performance
Assessing formation damage due to completion
Evaluation of DSTs and other well tests
Workover implementation
Indication of scale deposits
Design of artificial lift systems
Formations which have undergone pressure depletion may not withstand a static column of fresh water
Basic completion design and installation in any well under any balance condition.

Hydrostatic Pressure

The hydrostatic component is the pressure exerted by the column of liquid/gas in the wellbore. This pressure is a
function of the fluids present and consequently can vary, from a static bottomhole value if the well is shut-in (or
not flowing), to a hydrostatic pressure reflecting changing fluid ratios during flow. Under static conditions,
without flow, there are no frictional contributions and a static bottomhole pressure is calculated from a
summation of surface pressure and the hydrostatic head.

Calculating the static bottomhole pressure for a liquid-filled wellbore involves simply multiplying the fluid
gradient (density) times the true vertical depth.

For a vertical well filled with gas, the static condition is given by integrating the density over the depth interval.
The form of the integration is given by:

(31)

where:

Z = gas deviation factor (dimensionless). Z can be regarded as a term by which


pressure must be corrected to account for departure from the ideal gas equation
pV = ZnRT
p = absolute pressure (psia)
g = specific gravity of the gas (dimensionless, air = 1.0)
T = absolute temperature (R)
L = length (feet)

If the temperature and pressure are taken at an average value (indicated by the bar in the following equation),
the following equation accounts for gas compressibility:

(32)
where:
pws = static bottomhole pressure (psia)
pts = static wellhead pressure (psia)
H = true vertical depth (feet)

The temperature is commonly taken as the arithmetic mean of the bottomhole and wellhead temperatures,
assuming a linear variation with depth. is taken as the value at the arithmetic mean temperature and the
arithmetic mean pressure. Because of the interrelationships between parameters, this bottomhole pressure
equation is generally solved by trial-and-error. While you likely have a numerical model or spreadsheet to do
these calculations, it is instructive to understand some of the background. The iterative steps are:

1 Guess a reasonable value for pws.

2 Determine at and .
3 Calculate pws from Equation 32. This is an estimate of the bottomhole pressure. The hydrostatic component
equals the surface pressure minus the bottomhole pressure.
4 Presuming that the value of pws differs substantially from its original estimate, refine the assumed value
for pws and repeat the process until adequate agreement occurs.

If you use the method of Sukkar and Cornell, 1955, trial and error procedures are not required. For even more
precise calculations, you can use the method proposed by Cullender and Smith (1956). Good references on this
topic are Bradley, 1991 and Ikoku, 1980.

Friction

Using the methods in the previous section, you can estimate the hydrostatic head for a well containing liquid,
gas, or multiphase fluids during a completion operation. The next component to consider in calculating the
bottomhole pressure is the friction that exists when fluids are moving. The complex frictional forces of fluids
flowing in a wellbore depend on the type of fluid flow regime. This section considers frictional effects for:

Single Phase Liquid Flow


Single Phase Gas Flow
Multiphase Flow

Single Phase Liquid Flow


For liquid flow in the wellbore, the pressure drop per unit length (for friction effects only) is given by the
following equation:

(33)

where:
dP = pressure change (psi)
dL = length of the segment over which the pressure drop is calculated (feet)
p = 3
density (specific gravity x 62.4) (lbm/ft )
f = Moody friction factor (dimensionless)
v = velocity (ft/second)
D = equivalent diameter (inches)

The term equivalent diameter is used to discriminate between pipe flow and annular flow. For conventional pipe
flow, Equation 33 is used and D is the internal diameter of the pipe over the length being evaluated. For annular
flow, the effective diameter is:

(34)

For annular flow, use D = Douter Dinner in the second form of Equation 33. In the first form of Equation 33,
5 3 2
replace D with (Douter Dinner) x (Douter + Dinner) .

The length term in Equation 33 is based on measured depth because friction acts over the full length of the
string (unlike hydrostatic head, which is based on true vertical depth). If there are changes in string dimensions
and/or inclination, it is common practice to divide the well up into individual increments and sum the friction over
the individual discrete lengths. If the wellbore contains compressible fluids and you use fundamental friction
equations, you must discretize the wellbore (divide it into discrete lengths).

As reflected in Equation 33, during flow, irreversible energy losses occur. With the exception of completely
laminar flow, these energy losses cannot be predicted theoretically and are usually accounted for by using the
friction factor (as shown in Equation 33). For consistency, only the Moody friction factor will be used (the Moody
friction factor is four times the Fanning friction factor (fF)). The friction factor (f) and the relative roughness (e/D)
are related to the Reynolds Number. The Reynolds Number is calculated by:

(35)

where:
e = absolute roughness (feet)
D = diameter (feet)
S.G. = specific gravity (dimensionless)
v = velocity (feet/second)
NRe = Reynolds Number (dimensionless)
m = viscosity (1bm/ft-sec or cP)

The friction factor depends on the specific flow regime in the wellbore, as shown in Table 5. With the exception of
laminar flow, many empirical representations of the friction factor are available for turbulent and transitional flow
regimes. Most of these equations yield approximately the same results (see Figure 12 for a relative roughness of
0.0006).

Table 5. Summary of Various Friction Factor Formulas


Flow Regime Reynolds Number Moody Friction Factor Comments
Laminar < 2000
Flow Regime Reynolds Number Moody Friction Factor Comments
Independent of
roughness
26 2000 < NRe < 4000
Critical

26 4000 < NRe < (200D/e) Colebrook (1939)


Transition
1.16

26,28 NRe > 4000 Nikuradse


Turbulent

Figure 12. Comparison of Moody friction factors for various calculation methods for a relative
roughness of 0.0006. Woods curve is an explicit approximation of Colebrooks
relationship.

Though roughness is quite difficult to determine, particularly for tubulars that have been in service for some
time, some guidelines are available. Figure 13 shows an example. Using these guidelines, you can estimate
frictional pressure loss using Equation 33 and Table 5. It is usually appropriate to discretize the flow path and
solve over various increments, using the previous increment to provide an upstream boundary condition.
Commercial codes are available for these calculations.
Figure 13. Relative roughness of various pipes.

Single Phase Gas Flow

For steady-state flow, the energy balance is given by (notice that this contains the hydrostatic head component).

(36)

where:
r = 3
fluid density (lbm/ft )
p = pressure (psia)
v = average fluid velocity (ft/sec)
a = correction factor to compensate for variation of velocity over the tubing cross-
section, varying from 0.5 for laminar flow to 1.0 for fully developed turbulent flow.
A value of 0.90 is commonly used for practical gas flow problems.
H = distance in the vertical direction (feet)
f = Moody friction factor (dimensionless)
D = inside pipe diameter (feet)
L = length of the flow string (feet), L = H for a vertical section
= pressure drop due to kinetic effects,

= pressure drop due to friction,

ws = mechanical work done on or by the gas (ws = 0)


g = 2
acceleration due to gravity (ft/s )
gc = conversion factor (32.17 lbm - ft/1bf s )
2

Ignoring the kinetic energy term and integrating over an interval from point 1 to point 2 gives:
(37)
Combining using oilfield units and assuming constant temperature over an interval of interest:
(38)

where:
p = pressure (psia)
Q = gas flow rate (MMscfD)
= average temperature over the distance L, (R)

L=H = vertical distance (feet)


D = inside diameter of the pipe (inches)

Unfortunately, without assumptions or numerical solution, this equation cannot readily be integrated. Smith
(1950) proposed an approximate solution (based on the Weymouth equation):
(39)

where:
Q = volumetric flow rate measured at 14.656 psia and 60F (scfD)
T = effective average temperature (R)
Z = effective compressibility factor for the gas
d = pipe ID (inches)
p2 = sandface pressure at depth X (psia)
p1 = wellhead pressure (psia)
gg = gas gravity (dimensionless, air = 1.0)
X = difference in elevation between points 1 and 2 (feet)
f = Moody friction factor (dimensionless)

There are numerous other approximate or numerical formulations. Most of these are available in or superceded
by commercial software packages. For example:

Poettmann (1951) derived an expression accounting for the variation of the compressibility factor. It can be
applied to deviated wells.

You can use a modified version of the Sukkar and Cornell Method (see Katz et al., 1959) for straight or
inclined holes. Their relationship assumes steady-state, single phase flow, negligible kinetic energy, constant
temperature at some average value, and a constant friction factor over the length of the conduit.

Two-step approximations can be done using the Cullender and Smith method.

There are other relationships: the Panhandle formula; the Ford, Bacon and Davis formula; and the Clinedinst
flow equation.

Multiphase Flow
The procedures for evaluating multiphase flow are substantially more complicated than those for single-phase
flow. The level of complexity for multiphase flow is so high that numerical analysis and/or correlations are
commonly used for frictional evaluations. Some of the possible correlations that you can use for forecasting
pressure drop are:

Duns and Ros (1963) can be used for certain vertical flow situations.

Orkiszewski (1967) can be used for certain vertical flow situations. Orkiszewski outlined methods for
addressing flow of discrete bubbles, slug flow where the gas phase is more pronounced, a transition flow
(transitional from continuous liquid to continuous gas) and mist flow where the gas phase is continuous and
most of the liquid is present as entrained droplets. Computational versions of this method are widely used,
with some specific limitations.

Hagedorn and Brown, 1965, is probably recommended for most vertical flow situations. It was developed
from experimental measurements in an instrumented 1500-ft deep well. Neither liquid holdup or flow pattern
was measured but correlations were developed by assuming that the two-phase friction factor could be
obtained from the Moody diagram based on a two-phase Reynolds Number. Refer to Appendix A in Beggs
1991. This is recommended reading on this topic.

Beggs and Brill, 1973, deals with any inclination, two-phase flow. Liquid holdup and pressure gradients were
measured for flow of air and water in a laboratory environment, for inclinations between horizontal and
vertical. Liquid holdup is the fraction of an element of pipe that is occupied by liquid at some instant.

Mukherjee and Brill, 1983, developed a procedure for wells of any inclination with two-phase flow, and,

Dukler, 1959, provided correlations for horizontal flow.

The Hagedorn and Brown correlation commonly


gives good results for most situations.

There are some simpler methods as well. You can use the Poettmann-Carpenter method for vertical multi-phase
flow. Recognize that other methods, such as Hagedorn and Brown, are more (and sometimes much more)
accurate. Poettmann-Carpenter determined an f factor empirically and correlation charts were prepared for
various tubing sizes on which were plotted the total mass-producing rate QM versus the pressure gradient dP/dh.

Computational routines are commercially available (and should be used) for multiphase flow for high profile
situations. For rapid calculations, in addition to the Poettmann and Carpenter procedures, an extensive set of
gradient curves has been generated and published by Brown (1984). Gradient curves have been generated for
various tubing/annulus dimensions, rates and the specific fluid properties.

Restrictions
Inevitably, there will be pressure losses in in-line restrictions, including subsurface safety valves, surface or
bottomhole chokes, valves and fittings, etc., Beggs provides a clear discussion. Some of the key points are
summarized below.
Single Phase Gas Flow through Surface Chokes

Flow through a restriction can be critical or subcritical (sonic or subsonic). When the flow is critical, any
downstream disturbance will not affect the upstream pressure or the rate. Since chokes are intended to control
rate, they usually are designed for critical flow. A rule-of-thumb for distinguishing between critical and
subcritical flow states that if the ratio of downstream pressure to upstream pressure is less than or equal to 0.5,
then the flow will be critical. This is a closer approximation for single-phase gas than for two-phase flow.

(40)
where:
qsc = volumetric gas flow rate (Mscf/D)
d = I.D. of bore open to gas flow (inches)
gg = gas specific gravity (dimensionless, air = 1.0)
k = ratio of specific heats = Cp/Cv (dimensionless)
p1 = upstream pressure (psia)
p2 = downstream pressure (psia)
T1 = upstream temperature (R)
Z1 = compressibility factor at p1 and T1 (dimensionless)
CD = discharge coefficient (empirical, dimensionless)
Tsc = standard absolute temperature base (R)
psc = standard absolute pressure base (psia)
y = p2/p1
yc = critical pressure ratio (dimensionless)

When no information is available, the restriction is short with slightly rounded openings and critical flow
conditions exist, an approximation is (using a generic default value of CD = 0.82):

(41)

Single-Phase Liquid Flow through Surface Chokes

The following equation may be used (assuming subcritical flow, if necessary, a default value for the discharge
coefficient can be used, CD = 0.85):

(42)

where:
qL = liquid flow rate (STB/day)
d = choke diameter (inches)
p = psi
S.G. = liquid specific gravity

Two-Phase Flow through Surface Chokes

An empirical relationship in the critical regime is:

(43)
where:
pi = upstream pressure (psia)
qL = liquid flow rate (STB/day)
R = gas/liquid ratio (scf/STB)
d = choke diameter (inches)

Values of a, b and c have been proposed by different investigators. For very rough approximations, try using a
= 1.89, b = .00386 and c = 0.546. Sachdeva (1986) presented a more rigorous model.

Gas Flow Through a Subsurface Safety Valve

Flow will generally be subcritical - minimal downhole restriction is desirable. You can iterate and solve using the
API relationship:

(44)

where:
p1 = upstream pressure (psia)
p2 = downstream pressure (psia)
g = gas gravity (dimensionless, air = 1.0)
Z1 = gas compressibility at p1 and T1
T1 = upstream temperature (R)
qsc = gas flow rate (Mscf/D)
b = d/D
d = bean diameter (inches)
D = pipe inside diameter (inches)
CD = discharge coefficient (API suggests 0.9)
Y = expansion factor (dimensionless). The expansion factor determination is iterative
and may be calculated from Equation 44. Its value ranges between 0.67 and 1.0.
For quick estimates, a default value of 0.85 is often used.
K = the ratio of specific heats of the gas.

Two-Phase Flow Through a Subsurface Safety Valve

For two-phase flow through a subsurface safety valve, you can use:

(45)
where:
rn = 3
no-slip density (lbm/ft )
qL = "in-situ" total liquid rate
qg = "in-situ" gas rate
nm = mixture velocity through the choke (ft/sec)
p1 = upstream pressure (psi)
p2 = downstream pressure (psi)
CD = discharge coefficient
Nn = qg/qL = (1-lL)/lL
lL = qL/(qL+qg)
b = d/D
d = choke diameter
D = tubing inside diameter

All the fluid properties necessary for calculating the density and velocities are evaluated at upstream conditions
of pressure and temperature. p1 is iteratively determined from a known p2.

Pressure Losses Through Valves and Fittings

The energy losses associated with bends and other restrictions are considered to be supplemental to those
predicted by Bernoulli's basic equation and its standard representation using Moody's friction factor. To account
for these additional pressure losses, the concept of a loss coefficient is used. Typically, the loss coefficient is
indicated as K. In simplified terms the pressure loss for flow with bends and with other in-line obstructions can
be expressed as:

(46)

where:
d = diameter of pipe
f = friction factor for pipe flow
L = length of pipe
K = resistance coefficient depending on the type or size of fitting

An equivalent length, Le, can be calculated for each fitting by using the friction factor calculated for flow in the
pipe. All the Le values can then be added to the actual pipe length for the pressure drop calculation. K values
have been determined for single-phase flow and they are similar to values for two phase flow (for example, 0.15
for a gate valve, 0.2 -0.3 for elbows, 3.0 - 5.0 for globe valves, 6.0 - 8.0 for check valves). The pressure drops
will commonly be small enough to be ignored. A more basic fluid mechanics approach is provided by Granger,
1985. Granger indicated the loss coefficient for a rounded elbow could be determined from empirical
relationships.

(47)

If no values of r/a are available, use: k45 = 0.35 and k90 = 0.75. Most manufacturers can provide relevant
information.

The degree of pressure loss due to bending is generally taken to be a function of R (the radius of curvature and r
the inner radial dimension of the pipe. The term R/r is known as the relative radius. For large values of R/r,
such as those which would be associated with the transitional segment of a horizontal well, Rankin et al., 1989,
proposed a loss coefficient which is empirically correlated to the relative radius, a friction factor (f), bend angle in
degrees and a correction coefficient which depends on both the bend angle and R/r. These relationships were
empirically developed using data for water in fully developed turbulent flow and are therefore relevant to
conventional fluids (as opposed to lightened fluids). The transition from horizontal to vertical flow in the annulus
is particularly sensitive to momentum flux the tendency of the fluid to surge as head is overcome. This is
atttributable to the transition from a somewhat static pressure regime in the horizontal segment to a sudden
reduction in momentum as the fluid is forced to travel upward.

Additional pressure drops that can be considered are associated with tapered strings, tubing restrictions, the
wellhead and any horizontal lines or obstructions at the surface. Tapered strings can be evaluated numerically or
published gradient curves can be used. The gradient curves for the specific tubular or casing dimensions are
used in combination by manipulating the effective depth (using the bottom run depth of an upper tubing as the
top of the section below).

Various restrictions to flow can be present - e.g., safety valves, tubing nipples, downhole chokes. Some
manufacturer information is available. Similar information can be acquired for wellhead/BOPRBOP etc.
restrictions. Horizontal surface lines can be represented using one of the methods that have been described
above.

Using Tubing Intake Curves (Injection or Production)

Regardless of which method you use to estimate pressure drops, you will need to determine hydrostatic pressure
and friction at various flow rates. Using your selected method, you must generate tubing intake curves to
represent the flowing bottomhole pressure versus rate for the particular tubing/annulus situation. Your tubing
intake curves will be based on the specific set of parameters that you define (e.g., wellhead pressure, diameters,
lengths, fluid type, etc.). Once you have defined these parameters, you will construct a table of bottomhole
pressure versus rate. Finally, you will plot the tubing intake curves. An example is shown in Figure 14.
Figure 14. An example tubing intake curve.

Tubing intake curves depend on a number of environmental parameters. For example, different tubing intake
curves will exist for different wellhead pressures (refer to Figure 15).

Figure 15. Various tubing intake curves for different flowing wellhead pressures. As pwf
decreases, the wellhead pressure also decreases.

The tubing intake curve (flow rate that the tubing can accommodate) is also influenced by the size of the
tubulars (refer to Figure 16).
Figure 16. Schematic change in tubing intake curves for different size tubing.

You will recall that there are methods for inferring the drop in pressure for flow through the formation, and
through the completion. You can also see from the example tubing intake curves that there can be further
restrictions associated with flow in the wellbore. Because the formation, completion, and wellbore pressure
TM
drops are not independent, you must consider them using a systems approach. You can use Nodal Analysis
principles and techniques to evaluate the individual and combined effects of these parameters on well
performance.

Nodal Analysis
TM
Nodal Analysis was originally developed by Flopetrol Johnston. Nevertheless, the concepts are universally
applicable and used by all service and operating companies. In this method, you represent the reservoir-
completion-string-surface components by nodes that can be found in different parts of the well. A node is any
point at which a pressure drop is evidenced. Next, you evaluate the pressure drops at all points and determine
an equilibrium situation. Finally, you can determine the maximum potential injection.

The performance relationships indicate the flow rates that a formation can deliver and they can account for the
effect of a particular completion. The tubing intake curve accounts for the flow rate the string/annulus can
accommodate. You must superimpose the performance relationship and tubing relationships to determine rate
and pressure.

TM
A simple example of using Nodal Analysis or related concepts is shown in Figure 17. This schematic example is
for a hypothetical injector. The effect of the perforation density has been exaggerated for illustrative purposes. A
tubing intake curve developed for a particular combination of fluids is superimposed on the IPR curves. Using
these data and the bottomhole pressure range, injection rate is predicted. Expected wellhead pressures can be
determined by varying input into the calculation of the tubing intake curve.
Figure 17. Hypothetical injection well. Several performance curves are shown for various
perforating programs. Several tubing intake curves are superimposed for varying
wellhead pressures. Intersections indicate anticipated pressures and flow rates.

TM
You can use Nodal Analysis or equivalent methods to evaluate potential well performance. If these programs
are not available, do the appropriate calculations in a spreadsheet.

References
1. Good, P.A. and Kuchuk, F.J.: Inflow Performance of Horizontal Wells, paper SPE 21460 (1991).
2. Kuppe, F. and Settari, A.: A Practical Method for Determining the Productivity of Multi-Fractured
Horizontal Wells, CIM Edmonton 96, Edmonton, Canada (April28-May 2, 1996) and JCPT 1998.
3. Kaiser, T.M.V., Willson, S., and Venning, L.A.: Inflow Analysis and Optimization of Slotted Liners, paper
SPE 65517, 2000 SPE/PS-CIM Intl Conf on Horizontal Well Technology, Calgary, Canada, November 6-8.
4. Asadi, M. and Penny, G.S.: Large-scale Comparative Study of Pre-Packed Screen Cleanup Using Acid
and Enzyme Breaker, 2000 IADC/SPE Asia Pacific Oil and Gas Conf. Exhib, Brisbane, Australia, October
16-18.
5. Penberthy, Jr., W.L. and Shaughnessy, C.M.: Sand Control, SPE Monograph Series, Richardson, TX
(1992).
6. Behie, G.A. and Settari, A.: Perforation Design Models for Heterogeneous Multiphase Flow, paper SPE
25901 presented at the 1983 SPE Joint Rocky Mountain Regional/Low Permeability Reservoir Symposium,
Denver, CO, April 26-28.
7. Duns, H., Jr. and Ros, N.C.J.: Sixth World Petroleum Congress, Section 2, Paper No. 22, Frankfurt,
Germany (1963).
8. Orkiszewski, J.: JPT., 19, 829 (1967).
9. Hagedorn, A.R., and Brown, K.E.: JPT, 16, 203 (1964); 17, 475 (1965).
10. Beggs, D.H. and Bril, J.P.: A Study of Two-Phase Flow in Inclined Pipes, SPE 4007, JPT (1973).
11. Mukherjee, H. and Brill, J.P.: Liquid Holdup Correlations for Inclined Two-Phase Flow, SPE 10923, JPT
(1983).
12. Dukler, A.E.: Chem. Eng. Progress, 55(10), 62 (1959).
13. McLennan, J.D., Carden, R.S., Curry, D., Stone, C.R. and Wyman, R.E.: Underbalanced Drilling Manual,
Gas Research Institute, Chicago, IL (1997).
14. Brown, K.E.: The Technology of Artificial Lift, Volume 4, Pennwell, Tulsa, OK (1984).
15. Sachdeva, R., Schmidt, Z., Brill, J.P. and Blais, R.M.: Two-Phase Flow Through Chokes, SPE 15657,
st
paper presented at the 1986 (61 ) Annual Tech. Conf. Exhib., New Orleans, LA.
16. Granger, R.A.: Fluid Mechanics, Dover Publications, New York, NY (1995).
17. Rankin, M.D., Friessenhahn, T.J. and Price W.R.: Lightened Fluid Hydraulics and Inclined Boreholes,
SPE 18670, paper presented at 1989 SPE/IADC Drilling Conf., New Orleans, LA, February 28 March 3.
18. Supon, S.B. and Adewumi, M.A.: An Experimental Study of the Annulus Pressure Drop in a Simulated
Air-Drilling Operation, SPE Drill. Eng. (March 1991) 74-80.
19. Wang, Z., Rommetveit, R., Bijleveld, A., Maglione, R. and Gazaniol, D.: Underbalanced Drilling Requires
Downhole Pressure Management, Oil& Gas J. (June 16, 1997) 54-60.

You might also like