You are on page 1of 10

A Single-Phase Wellbore-Flow Model for

Horizontal, Vertical, and Slanted Wells


Liang-Biao Ouyang, SPE, Sepehr Arbabi, SPE, and Khalid Aziz, SPE, Stanford U.

Summary

In this paper, a single-phase wellbore-flow model is presented that


incorporates not only frictional, accelerational, and gravitational
pressure drops, but also the pressure drop caused by inflow. The
new model is readily applicable to different wellbore-perforation
patterns and well completions, and can be easily incorporated into
reservoir simulators or analytical, reservoir-inflow models.
It is found that the influence of either inflow or outflow depends
on the flow regime present in the wellbore. For laminar wellbore
flow, wall friction increases because of inflow but decreases because of outflow. For turbulent wellbore flow, inflow reduces the
wall friction, whereas outflow increases the wall friction. New
wall-friction-factor correlations for wellbore flows have been developed that can be applied to determine the wall-friction shear
and the frictional pressure drop for either inflow (production well)
or outflow (injection well) and for either laminar or turbulent
flow regimes.
Calculation results show that the accelerational pressure drop
may or may not be important compared with the frictional component, depending on the specific pipe geometry, fluid properties,
and flow conditions. It is recommended that the new wellbore-flow
model be included in wellbore/reservoir coupling models to achieve
more accurate predictions of pressure drop and inflow distribution
along the wellbore, as well as the well production or injection rates.
Introduction

Over the last decade, horizontal wells have become a well-established technology for the recovery of oil and gas. A considerable
amount of analytical and experimental work has been published on
various aspects of horizontal-well production, including transientflow and stabilized-inflow models, productivity indices, and coning
and cresting behavior. Although these methods provide insight into
the behaviors of horizontal wells, only a few of them consider
pressure drop along the wellbore and essentially infinite conductivity is assumed. In 1990, Dikken1 proposed the first semianalytical model to evaluate the production performance of a horizontal
well with the consideration of the wellbore-pressure drop resulting
from turbulent flow. Since then, others2-9 have presented different
coupling models for wellbore flow and reservoir inflow through
perforations. However, even in cases where pressure drop along a
wellbore is considered, only the frictional component is included
under most circumstances; pressure drop because of acceleration
and other effects is neglected. We show that, because of the
existence of perforation inflow, accelerational pressure drop can be
important relative to the frictional component and can significantly
influence the well-flow rates under some flow conditions. Furthermore, traditional methods to determine frictional pressure drop in
pipe flow that do not account for inflow are used in most of the
coupling models, which is not justified, because inflow changes the
wall friction in the wellbore.
In this paper, a single-phase wellbore-flow model is presented
that incorporates not only frictional, accelerational, and gravitational pressure drops, but also the pressure drop caused by inflow.
The new model is readily applicable to different wellbore-perforation patterns and well completions and can be easily incorporated
into reservoir simulators or analytical reservoir-inflow models. To
obtain the precise contribution of the frictional pressure drop along
Copyright 1998 Society of Petroleum Engineers
Original SPE manuscript received for review 6 October 1996. Revised manuscript
received 19 December 1997. Paper peer approved 13 January 1998. Paper (SPE
36608) first presented at the 1996 Annual Technical Conference and Exhibition held
in Denver, Colorado, 69 October.

124

the wellbore, new wall-friction-factor correlations for pipe flow


with perforation influx are also developed on the basis of the
published research work and our own horizontal wellbore experimental data. They can be applied in wellbore flow to determine
wall-friction shear and frictional pressure drop for either inflow
(production well) or outflow (injection well) and for either laminar
or turbulent flow.
The new wellbore-flow model has been used to study the relative
importance of frictional and accelerational components of pressure
drop along a horizontal well. Other sensitivity analyses have also
been conducted. Moreover, the new wellbore model has been
coupled with a simple reservoir model to study the performance of
different horizontal wells.
Literature Review

As early as 1904,10 mechanical engineers first investigated fluid


flow in channels or pipes with fluid injection or extraction through
wall surfaces. In the beginning, researchers focused only on flow
past a flat surface with injection or suction. The study of this aspect
is of great practical importance for the aerospace industry, and it has
enhanced the design of airplanes and improved their performance.
For example, only a little suction on the surface of an airfoil was
found to increase lift and decrease drag significantly, as well as alter
the heat-transfer behavior. This discovery has been applied in the
aerospace and other related industries, such as the chemical industry (membrane filtration,11 transpiration cooling12), and nuclear
industry (235U enrichment13). Surprisingly, more than 50 years had
passed since the work on flat plates before Yuan and Finkelstein14
investigated the laminar fluid flow in pipes in the presence of fluid
injection or extraction through the porous wall.
Most theoretical studies on wall-friction factor and heat-transfer
characteristics for laminar pipe flow with mass transfer through its
porous wall are based on similarity solutions in which the velocity
and temperature profiles are only functions of the radial coordinate
(they remain unchanged in the axial direction). Similarity solutions
have been found to exist in a certain range of the wall Reynolds
number NRe,w (injection-Reynolds number for the injection case and
extraction- or suction-Reynolds number for the suction case) but
not for other values.
Yuan and Finkelstein14 were the first to investigate the effect of
uniform injection and suction through the pipe wall on the twodimensional steady-state laminar fluid flow in a porous pipe by
solving the Navier-Stokes equations in cylindrical coordinates for
both very large and very small wall Reynolds numbers. Kinney15
examined the frictional and heat-transfer charateristics of fully
developed laminar flow in porous pipe by means of numerical
solutions of the momentum and energy equations. The fluid flow
in a straight, circular tube with uniform inflow or outflow was
considered. The corresponding Navier-Stokes equations were reduced to an ordinary differential form by applying similarity
analysis. Olson and Eckert16 performed experimental studies of
turbulent air flow in a porous circular pipe with uniform air
injection through the pipe wall. The fully developed turbulent air
flow, at Reynolds numbers of 28,000 to 82,000, entered the pipe
while air was injected uniformly through the wall at various ratios
of injection velocity to the average velocity at the entrance, ranging
from 0.00246 to 0.0584. Kinney and Sparrow17 presented results
for frictional, heat-transfer, and mass-transfer characteristics of
fully developed turbulent flow in pipes with suction. They assumed
that the velocity and temperature fields were locally self-similar.
The variation of the friction factor with the relative suction velocity,
veq/v, was given in graphical form for parametric values of the
SPE Journal, June 1998

Reynolds number. Merkine et al.18 extended these results to account for the effect of suction on the level of turbulence in the fluid
(suction decreases turbulence level). Their results show good agreement with the measurements of velocity profiles by Weissberg and
Berman.19 Unfortunately, no results of the friction factor were
given. By modifying Kinney and Sparrows analysis, Doshi and
Gill20 obtained an excellent agreement with the experimental velocity profiles of Weissberg and Berman. Lombardi et al.21 investigated the fully developed turbulent flow along a pipe with air
injection through its porous wall and found that the Nusselt number
in the downstream region of the pipe could be related to local
parameters, such as the local axial-flow Reynolds number and the
injection Reynolds number, without additional dependence upon
the inlet-axial Reynolds number and axial position.
It is quite interesting to note that the characterisics of pipe flow
with wall mass transfer are different from those of channel flow or
flow past a flat plate. For example, considering the laminar flow
case, the local friction factor increases with an increase of wall
Reynolds number for pipe flow but decreases for channel flow.
Moreover, the suction-induced transition to turbulent flow (suctioninduced instability) and the existence of multiple solutions for pipe
flow have no counterpart in channel flow.22
The different behavior of the local friction-factor change with
inflow for laminar flow and for turbulent flow is another interesting
fact about pipe flow. For laminar flow, the local friction factor
increases with an increase in the injection Reynolds number,
whereas it decreases for turbulent flow. The underlying mechanisms will be discussed later in this paper.
Fluid flow in pipes with wall mass transfer did not interest many
petroleum engineers until the horizontal-well technology was introduced and widely applied in the petroleum industry starting in
the 1980s. For fluid flow in horizontal wells, the flow format
is quite similar to pipe flow with mass transfer through its porous wall. The main differences between these two types of flow
are as follows.
In horizontal wells, the mass transfer is normally through
perforations, whereas, in the case of pipe flow, it is through pores
in the wall. In other words, the effective perforation density is very
large (theoretically infinite) for the porous-pipe flow case. Nevertheless, in the case of openhole completions, the horizontal-well
and porous-pipe flow problems are conceptually identical.
The injection rates are usually quite small in the case of
porous-pipe flow; this is not necessarily the case for wellbore flow.
When there is no mass transfer through the pipe wall, the
effective pipe roughness may be very different from the actual pipe
roughness in a horizontal well because of the effect of perforations
on the axial flow (such as flow separation, cavity flow, or secondary
flow), but it changes only slightly from the actual value for the
porous-pipe flow case.
Little information is available with regard to the significance of
the above-mentioned differences on the prediction of horizontalwell behavior. Although much research work has been completed
for the porous-pipe flow problem, these results may not be directly
applicable to horizontal wells. Recognizing this fact, petroleum
engineers began to study the horizontal-well flow problem from
the late 1980s.
Kloster23 studied flow resistance in a perforated pipe, both with
and without flow injection through the pipe wall, by conducting
experiments on a pipe of 656 in. outside diameter and 17 ft in length.
Asheim et al.24 stated that the total pressure drop along a perforated
pipe is made up of wall friction and inflow acceleration and
computed the wall friction factor in the same way as for a regular,
unperforated pipe. Ihara et al.25 studied channel flow with continuous influx into the horizontal channel from an oil-reservoir model.
They stated that pressure gradients increase almost uniformly in the
test channel because of the confluence of influx and axial flow, and
the resulting pressure drop increases linearly with influx velocity.
Because of perforations, effective pipe roughness may be different from the original roughness even without inflow or outflow
through perforations. Su and Gudmundsson26 measured the mass
flow rates and water-column heights for water flow along a vertical
pipe. Data for flow, both with and without perforations, were
SPE Journal, June 1998

Fig. 1Mass and momentum change along a wellbore with


mass transfer through perforations.

collected. They introduced a roughness function in the frictionfactor calculation to account for the effect of perforations on pipe
roughness and correlated the roughness function as seven times the
perforation-to-casing-diameter ratio, based on their measurement.
For fluid flow in a pipe with one single perforation, Yuan et al.27
developed a correlation for the apparent friction factor by matching
their experimental data. (Apparent friction factor is equivalent to
dimensionless total pressure drop along a pipe section. It includes
different pressure-drop components, such as frictional, accelerational, and gravitational. Therefore, the apparent friction factor is
not exclusively related to wall friction. All the new correlations
presented in this paper are for the Fanning friction factor, which is
defined as the dimensionless pressure drop caused solely by wall
friction.) However, there is significant difference between wellbore
flow and pipe flow with one single perforation. Moreover, pipe
flow with one single perforation is not applicable to real horizontal
wells; hence, it is questionable whether their correlation can be used
for wellbore flow. Recognizing this deficiency, they extended their
experiments to pipe flow with more perforations and developed a
new correlation for the apparent friction factor.27
Although single-phase pipe flow with wall mass transfer has
been the topic of numerous research activities, more work is
necessary, considering the following observations.
No general correlation exists for determining the wallfriction factors for fluid flow in a wellbore with inflow or outflow
through perforations.
The accelerational- and inflow-directional-pressure drops are
neglected in most wellbore-flow models or wellbore/reservoir coupling models.
The wall-friction shears are usually evaluated with frictionfactor correlations for pipe flow without wall mass transfer. Therefore, the impact of mass transfer through the pipe wall is not included.
General Wellbore Flow Model

Consider fluid flow in a wellbore as shown in Fig. 1 and assume


single-phase flow of an incompressible Newtonian fluid under
isothermal conditions with no heat transfer to and from the fluid to
the environment. Furthermore, assume that no mechanical work is
done on or by the fluid during its passage through the pipe (no shaft
work or work of compression). With these assumptions, the momentum-balance equation takes the form28
@~pA!2 2 ~pA!1 #
5

nDx
g
rAI vr nx 2 rA# Dx sin u,
BI gc
gc

1
1
rA v2 2
rA n2 2tw S Dx
B1 gc 1 1 B2 gc 2 2
. . . . . . . . . . . . . . . . . . . (1)

which can be rearranged to get the pressure-gradient equation,

SD

r d n2
S
g
nr AI
dp
2 tw 1
5 2
n n 2 r sin u. . . . . . (2)
dx
gc dx B
A aI gc A r x
gc
Eq. 2 indicates that the overall pressure gradient consists of four
different components.
125

The pressure gradient caused by kinetic-energy change (accelerational effects). This term should be zero for incompressible
fluid flow in pipes with constant inside diameter (ID) and without
inflow or outflow through the pipe wall. Obviously, it will not be
zero for wellbore flow with wall-mass transfer.
The frictional-pressure gradient, which depends on both axial
and perforation flows.
The pressure gradient caused by inflow direction, which is
called inflow-directional-pressure gradient in this paper. It may
help or hinder the axial flow depending on the inflow direction.
The gravitational pressure gradient. It is reasonable to assume
that the gravitational-pressure gradient is trivial and thus negligible
for horizontal wellbore flow.
To describe quantitatively the relative importance of different
pressure-gradient components, the following three dimensionless
numbers are introduced: Raf 5 the ratio of the accelerational and
frictional pressure gradients; Rgf 5 the ratio of the gravitational and
frictional pressure gradients; and Rda 5 the ratio of the directional
and accelerational pressure gradients. With these definitions, the
total pressure gradient takes the simple form,
dp
4tw
5 2
@1 1 Raf ~1 2 Rda ! 1 Rgf #. . . . . . . . . . . . . . . . . . (3)
dx
d
The closer the wellbore location is to the horizontal-well, toe
(x 5 0), the smaller the local wellbore-production rate qw; thus, the
larger the Raf and the more important the accelerational pressure
gradient. In contrast, near the heel of the horizontal well (x 5 L),
the local production rate qw becomes large and close to the total
well-production rate, whereas the specific influx (inflow rate/
wellbore length) qe does not change significantly, so Raf is small,
and the accelerational-pressure gradient is small and may be negligible.
For the uniform inflow case, it can be shown that28:
In the laminar-flow regime, Raf is dependent only on fluid
properties, inflow rate, and pipe ID, but it is independent of location
x and pipe roughness . The larger the qe, the larger the Raf, but the
shorter the laminar-flow length Lwf.
In the turbulent-flow regime, the friction factor, f, depends on
the local Reynolds number, the wall Reynolds number, and the
relative pipe roughness; consequently, the Raf depends on location
x, pipe geometry (pipe ID and pipe roughness), fluid properties, and
inflow rates.
Because turbulent flow occurs along almost the whole wellbore
section for most practical situations, it is anticipated that the
momentum-correction factor B does not change much for different
velocity profiles29; therefore, B can be taken as constant. Besides,
the derivative, dv/dx, can be easily obtained from the mass-balance
equation. Hence, Eq. 2 can be rewritten as

tw 5

g
d dp 2nr AI
nr AI 2
2 2
nn 1
n sin 2g 2 r sin u .
4 dx B gc A I 2aI gc A I
gc
. . . . . . . . . . . . . . . . . . . . . . . . . . (4)

On the basis of Eq. 4, the wall friction shear stress and, hence,
the wall friction factor, can be calculated. The wall friction factor
is expected to depend on the local axial Reynolds number, the
injection Reynolds number, the effective relative pipe roughness,
and, possibly, the inflow/axial flow rate ratios.
Effects of Wall Mass Transfer on Friction

For fluid flow in horizontal wells, inflow or outflow velocity is


relatively low compared with axial velocity along most of the
wellbore. Under these circumstances, wall mass transfer has the
following characteristics.
Laminar Flow. When mass transfer through the pipe wall exists,
the parabolic velocity profile is inappropriate for describing the
velocity distribution over the pipe cross section. The velocity
should increase for inflow but decrease for outflow, following the
law of mass conservation. Although inflow leads to the increase of
126

axial velocities across the whole pipe section, as can be imagined,


the axial velocities near the pipe wall will increase more than those
closer to the pipe axial centerline. Similarly, outflow decreases
axial velocities near the wall more significantly than those away
from the wall. As a result, the velocity gradient on and near the pipe
wall increases because of inflow but decreases because of outflow.
Correspondingly, the wall friction increases for the inflow case but
decreases for the outflow case.
Turbulent Flow. With the presence of mass transfer through
perforations, the time-averaged velocity profile for turbulent pipe
flow is altered because of the interaction of axial flow and the wall
inflow or outflow. Inflow lifts and expands the turbulent boundary
layer and, thus, increases the axial velocity beyond the layer.
However, it decreases the velocity within the layer to follow the
mass conservation law. As a consequence, the axial velocity gradient near the pipe wall decreases, and so does the wall friction
shear stress. On the contrary, outflow lowers and reduces the
boundary layer and, thus decreases the average velocity outside the
layer. However, it increases the velocity inside the layer and results
in an increase of the axial velocity gradient near the pipe wall and
hence the wall friction shear stress. This analysis is consistent with
the numerical observations of Kinney and Sparrow17 for pipe flow
with outflow through the pipe wall.
New Wall-Friction-Factor Correlations

As discussed earlier, mass transfer through the pipe wall affects the
wall-friction shear. The influence of either inflow or outflow
depends on the flow regime present in the wellbore. The inflow
(production well) increases the wall friction for laminar flow while
decreasing it for turbulent flow. In contrast, outflow (injection well)
decreases the wall friction for laminar flow while increasing it for
turbulent flow. In other words, the wall friction is different from
that of pipe flow with no inflow or outflow. Therefore, frictionfactor correlations for pipe flow without inflow or outflow cannot
be used for wellbore flow with both axial flow in the pipe and
inflow or outflow through perforations.
Laminar Flow. Kinney15 numerically solved the problem of fully
developed laminar flow in a porous pipe where both axial flow and
inflow or outflow through the pipe wall were present. The results
show that the ratio between the local friction factor, f, and the
no-wall-flow friction factor, f0 (which is defined, in this paper, as
the friction factor computed from the correlations for pipe flow with
no mass transfer through the pipe wall by use of local Reynolds
number and effective pipe roughness), is dependent only on the wall
Reynolds number. Unfortunately, this relationship was given in a
graphical form and no equation was supplied. Hence, it cannot be
used easily in a wellbore-flow model.
Figs. 2 and 3 show the calculated ratios between the local friction
factor and the no-wall-flow friction factor corresponding to different wall Reynolds numbers NRe,w, based on the reduced ordinary
differential equation for the dimensionless stream function and the
numerical procedure provided in Kinney.15 Fig. 2 also shows the
prediction of the local wall friction factor by the Yuan and Finkelstein equation.14 Because the Yuan and Finkelstein equation was
obtained by means of a perturbation method in which the injection
wall Reynolds number, NRe,w, was assumed to be sufficiently small,
it is anticipated that this equation can only be used for small NRe,w
(say NRe,w , 2.0).
These data have also been used to develop the following frictionfactor correlations by means of a fast and effective nonlinear
regression procedure called the Polytope method.30, 31
For inflow (production well) (NRe,w . 0),
f5

16
0.6142
~1 1 0.04304 NRe,w
!.
NRe

. . . . . . . . . . . . . . . . . . . . . . . (5)

For outflow (injection well) (NRe,w , 0),


f5

~2NRe, w !1.3056
16
1 2 0.0625
.
NRe
~NRe, w 1 4.626!20.2724

. . . . . . . . . . (6)

SPE Journal, June 1998

by use of Olson and Eckerts experimental data16 for turbulent air


flow in a porous pipe with uniform air injection through the pipe
wall. The new correlation is of the form

S D G

f 5 f0 1 2 29.03

NRe, w
NRe

0.8003

. . . . . . . . . . . . . . . . . . . . . . . (7)

As shown in Fig. 4, the new correlation provides satisfactory


predictions of wall friction factor for porous-pipe flow with inflow
through the pipe wall. The no-wall-flow friction factor, f0, can be
determined from the Colebrook-White equation or from one of its
explicit approximations.29
For turbulent pipe flow with outflow through the porous-pipe
wall, the local friction factor can be calculated by the Wallis
correlation32 based on Kinney and Sparrow17 data.

f 5 f0 1 2 17.5

NRe,w
.
N0.75
Re

. . . . . . . . . . . . . . . . . . . . . . . . . . . . (8)

Both correlations (Eqs. 7 and 8) are obtained on the basis of


experimental data for porous-pipe flow. It is unclear whether they
can also be used for normal wellbore-flow conditions. We have
applied Eq. 7 to analyze the single-phase wellbore-flow data in the
1995 Stanford Horizontal Wellbore Experiments33 and found that
Eq. 7 overpredicts the wall friction factor. Furthermore, it was
found that the ratio between the local friction factor and the
no-wall-flow friction factor does not depend on the wall Reynolds
number to axial Reynolds number ratio; instead, it depends only on
the wall Reynolds number. Therefore, a new correlation for the
local friction factor was developed.
0.3978
f 5 f0 @1 2 0.0153 NRe,
. . . . . . . . . . . . . . . . . . . . . . . . . . (9)
w #.

Fig. 2Effect of inflow on wall friction factor (laminar flow).

As shown in Fig. 5, Eq. 9 is a satisfactory correlation for local wall


friction factor for single-phase turbulent wellbore flow.
Results and Discussion

Accelerational- to Frictional-Pressure-Gradient Ratio, Raf .


Figs. 6 through 8 show how the dimensionless number, Raf , the
accelerational- to frictional-pressure-gradient ratio, changes with

Fig. 3Effect of outflow on wall friction factor (laminar flow).

Figs. 2 and 3 also show comparisons between friction factors


predicted by the new correlations and those of published
numerical experiments.
Turbulent Flow. A new correlation for the local wall friction
factor for turbulent flow has been developed in the present paper
SPE Journal, June 1998

Fig. 4 Wall-friction-factor correlation for turbulent pipe flow


with inflow.
127

Fig. 5Wall-friction-factor correlation for turbulent wellbore flow.

Fig. 7Well-production-rate influence on Raf.

Fig. 6 Wellbore length influence on Raf.

Fig. 8 Oil-viscosity influence on Raf.

wellbore location for uniform-influx horizontal wells with different


wellbore lengths, production rates, and oil viscosities. These figures
show that three regions exist along the wellbore: the laminar-flow
region, the partially developed turbulent-flow region, and the fully
developed turbulent-flow region. The dimensionless number, Raf ,
is a constant in the laminar-flow region that depends on fluid
properties (viscosity and density), production rate, and pipe geometry (pipe ID and pipe roughness). If all other parameters are fixed,
then the shorter the horizontal well, or the higher the production
rate, or the less viscous the oil, the larger the value of Raf , but the

shorter the length of the laminar-flow region. For the fully developed turbulent-flow region, all the Raf curves merge into one, which
varies only with wellbore location. Between the laminar and the
fully developed turbulent-flow region lies the partially developed
turbulent-flow region, where Raf varies with wellbore location, fluid
properties, production rate, and pipe geometry.

128

Analysis of Experimental Data. The new wellbore model has


been applied to analyze single-phase data from the 1995 Stanford
Horizontal Wellbore Experiments.33 Figs. 9 and 10 show variations
SPE Journal, June 1998

evaluated from the Colebrook-White equation with local Reynolds


number and effective pipe roughness. Fig. 9 indicates that the
apparent friction factor is larger than the no-wall-flow friction
factor, whereas the local friction factor is smaller than the no-wallflow friction factor. The apparent friction factor is augmented
because of the accelerational-pressure drop induced by inflow, and
the local friction factor is reduced (not increased) by inflow. This
observation is consistent with our analysis of inflow, discussed in
Effects of Wall Mass Transfer on Friction. Fig. 10 shows that
accelerational-pressure drop is large enough to be as important as
the frictional component for this case. Therefore, the result will not
be correct for this case if the accelerational-pressure drop is
ignored. For example, if the accelerational-pressure drop is not
included, then the local wall friction factor thus obtained is the
apparent friction factor as shown in Fig. 9, which is larger than the
no-wall-flow friction factor. This is contradictory to both theoretical analyses and experimental observations.

Fig. 9 Comparison of different friction factors.

Partially Perforated Horizontal-Well Example. The partially


perforated horizontal well is one in which only a part of the
wellbore is perforated for inflow. Also, it is assumed that the
wellbore-flow rate before the perforation point, q0, may or may not
be zero.
Figs. 11 and 12 show pressure gradients and cumulative pressure
drops from the start of the perforation in a partially perforated
horizontal well (Well 1 in Table 1). Also shown in Fig. 11 are the
Raf and the frictional-pressure gradient determined from the nowall-flow friction factor, f0. There is nontrivial difference between
the actual frictional-pressure gradient and the frictional-pressure
gradient obtained from the no-wall-flow friction factor. In this case
the accelerational-pressure gradient is comparable with the frictional component, and the dimensionless number Raf changes from
1.27 at the starting point of the well perforations to 0.71 at the end
of the perforations. The cumulative accelerational-pressure drop
from the start of the perforations is about the same as the cumulative
frictional pressure drop (Fig. 12).
Fully Perforated Horizontal-Well Example. In contrast to partially perforated horizontal wells, fully perforated horizontal wells
have perforations over the entire wellbore length.

Fig. 10 Comparison of different pressure-drop components for


IA2-6 experiment.32

of different friction factors and pressure drops over each 10-ft


section with wellbore location for Series IA2-6 experiments33 (the
axial water-flow rate and the perforation-flow rate are 10,300 and
6,900 B/D, respectively). Three friction factors are considered: the
total or apparent friction factor, fT, obtained on the basis of measured pressure drop; the local friction factor, f, obtained from
experimental data by means of the new wellbore-flow model (to
remove the accelerational and inflow-directional components from
the total pressure drop); and the no-wall-flow friction factor, f0,
SPE Journal, June 1998

Fig. 11Variation of pressure gradient with wellbore location


(Well No. 1).
129

Fig. 13Variation of pressure gradient with wellbore location


(Well No. 2).
Fig. 12Variation of cumulative pressure drop with wellbore
location (Well No. 1).

TABLE 1HORIZONTAL WELL PARAMETERS


Parameter

Well 1

Well 2

Well 3

Wellbore length, ft
Perforation length, ft
Pipe ID, in
Relative pipe roughness
Perforation ID, in
Perforation density,
shots/ft
Inflow direction, degrees
Fluid density, lbm/ft3
Fluid viscosity, cp
Wellbore rate at
perforation start point,
B/D
Inflow rate, B/D
Specific production
index, B-D/ft-psi

1,000
100
6.18
0.0002
0.18
8

1,870
1,870
4.5
0.0001
0.18
10

1,000
1,000
6.0
0.0002
0.18
10

90
62.4
0.878
7,000

45
52.4
2.5
0

45
62.4
1.0
0

7,000
N/A

7,500
N/A

N/A
2.0

Fig. 13 shows the change in pressure gradients with wellbore


location for Well 2 in Table 1. In this case, the dimensionless
pressure gradient ratio, Raf , ranges from 0.8 to 0.05, and the
cumulative accelerational-pressure drop is only about 6% of the
total pressure drop along the whole wellbore. In this case, the
accelerational-pressure gradient is quite small compared with the
frictional-pressure gradient, and only minor errors will result if it
is neglected.
Coupling the New Wellbore Model With a Reservoir-Inflow
Model. It is assumed in the previous discussion that the specific
inflow rate (i.e., the inflow rate/wellbore length) is the same
everywhere along the whole wellbore. This is not necessarily true
in a real horizontal well because the wellbore pressure near the well
toe is higher than that at the well heel because of the wellbore130

pressure drop. The pressure drawdown near the well toe is less than
that near the well heel; therefore, the specific inflow rate near the
well toe should be smaller than that near the well heel. On the one
hand, wellbore pressure, or pressure drop, is needed to determine
the specific inflow-rate distribution along the wellbore; on the other
hand, the specific inflow-rate distribution is an indispensable parameter to calculate wellbore pressure, or pressure drop. Hence, a
model that incorporates both the wellbore-fluid flow and the fluid
flow from reservoir into wellbore is required for determining
specific inflow-rate distribution, wellbore-pressure drop, and overall well-production rate.
Because the reservoir-inflow model is not the main concern of
this paper, a simple model used by Dikken1 and Novy9 is applied
to describe the reservoir-fluid flow.
qe ~x! 5 Js @pe 2 pw ~x!#,

. . . . . . . . . . . . . . . . . . . . . . . . . . . (10)

where the specific productivity index, Js, is assumed to be constant.


As can be found from the discussion in General Wellbore Flow
Model, the total pressure gradient for horizontal-wellbore flows
consists of three components: frictional, accelerational, and inflowdirectional. For Well 3 in Table 1, the frictional component is the
largest part of the total pressure gradient; the accelerational component is in the range of 20 to 80% of the total pressure gradient,
depending on the wellbore location (Raf changes from about 7.0 to
0.25); and the inflow-directional component contributes very little
and can be neglected (Fig. 14). Correspondingly, the cumulative
pressure drop along the wellbore is caused primarily by wall
friction (which is also affected by inflow), whereas acceleration or
kinetic-energy changes only contribute about 5 to 20% to the
overall pressure drop (Fig. 15).
Indeed, the specific inflow rate is not constant along the wellbore
because of the wellbore-pressure drop (Fig. 16). Significant differences occur between the real specific-inflow rate predicted by
considering pressure drop along the wellbore and the fictitious
specific-inflow rate for which no pressure drop along the wellbore
is assumed. For a more accurate prediction of the specific-inflowrate distribution, the accelerational-pressure drop should also be
included in the wellbore model; otherwise, approximately 10%
overestimation of the specific inflow rate is likely to occur.
It can be assumed that the well-production rate is proportional to
the wellbore length if no pressure drop occurs along the wellbore.
This statement is invalid for many real horizontal wells. Fig. 17
SPE Journal, June 1998

Fig. 14 Variation of pressure gradient with wellbore location


(Well No. 3).
Fig. 16 Specific inflow-rate distribution (Well No. 3).

Fig. 15Variation of cumulative pressure drop with location


(Well No. 3).

shows the production-rate change with wellbore length. Wellproduction rate decreases significantly because of wellbore-pressure drops and is likely to reach an asymptotic value at a certain
wellbore length. Beyond this, no additional production will be
contributed by additional well length.
Conclusions

A general, single-phase, wellbore-flow model is presented in this


paper. The new wellbore model is readily applicable to different
SPE Journal, June 1998

Fig. 17Variation of overall well-production rate with wellbore


length (Well No. 3).

wellbore-perforation patterns and well completions and can be


incorporated easily into reservoir simulators or analytical reservoirinflow models. Inflow and outflow affect wall friction shear. The
influence of either inflow or outflow depends on the flow regime
present in the wellbore. Inflow (production wells) increases the wall
friction for laminar flow but decreases the friction for turbulent
flow. The opposite is true for outflow (injection wells).
New wall-friction-factor correlations for pipe flow with wall
mass transfer are presented and can be applied to determine
frictional-pressure drop for either inflow (production well) or
131

outflow (injection well) and for either laminar flow or turbulent


flow.
It is shown that accelerational-pressure drop may or may not be
important compared with the frictional component, depending on
the specific pipe geometry, fluid properties, and flow conditions.
Nomenclature
A 5 pipe cross-sectional area, L2, ft2.
AI 5 cross-sectional area in each perforation, L2, ft2
A# 5 average pipe cross-sectional area, L2, ft2
B 5 momentum correction factor
d 5 pipe diameter, L, ft
f 5 Fanning friction factor
f0 5 no-wall-flow friction factor
fT 5 total or apparent friction factor
g 5 acceleration because of gravity, L/t2, ft/sec2
gc 5 conversion factor
Js 5 specific productivity index, L3t/m, STB-D/(psi-ft)
Lwf 5 wellbore laminar-flow length, L, ft
n 5 perforation density, 1/L, ft21
NRe 5 Reynolds number
NRe,w 5 wall Reynolds number, based on pipe inner diameter and
equivalent inflow/outflow velocity, veq
p 5 pressure, m/Lt2, psi
pe 5 pressure at boundary of well drainage volume, m/Lt2, psi
pw 5 wellbore pressure, m/Lt2, psi
qe 5 wall inflow/outflow rate per unit wellbore length, L2/t,
STB/(D-ft) or scf/(D-ft)
qw 5 local wellbore-flow rate, L3/t, STB/D or scf/D
ratio of the accelerational and the frictional pressure graRaf
dients
Rda 5 ratio of the directional and the accelerational pressure
gradients
Rgf 5 ratio of the gravitational and the frictional pressure gradients
S 5 wellbore perimeter, L, ft
v 5 velocity averaged over a cross section, L/t, ft/sec
veq 5 equivalent inflow/outflow velocity, qe /(pd) L/t, ft/sec
vr 5 radial component of inflow/outflow velocity, L/t, ft/sec
vt 5 inflow/outflow velocity in each perforation, L/t, ft/sec
vx 5 axial component of inflow/outflow velocity, L/t, ft/sec
x 5 axial coordinate, L, ft
xD 5 dimensionless axial distance away from inlet16
g 5 inflow direction angle
5 absolute pipe roughness, L, ft
u 5 wellbore inclination angle from horizontal
r 5 fluid density, m/L3, lbm/ft3
tw 5 wall-friction shear, m/Lt2, lbf/ft2
Acknowledgments

The authors thank Donald Schroeder of Marathon Oil Co. for


designing the equipment and supervising the Stanford Horizontal
Wellbore Experiments at Marathon. Thanks also go to the Stanford
U. Petroleum Research Inst.s Reservoir Simulation Industrial
Affiliates Program (SUPRI-B) and the Stanford Project on the
Productivity and Injectivity of Horizontal Wells (SUPRI-HW) for
providing the financial support for this study. This work was
partially supported by the U.S. Dept. of Energy contract DE-FG22
93BC14862.
REFERENCES
1. Dikken, B.J.: Pressure Drop in Horizontal Wells and Its Effect on
Production Performance, JPT (November 1990) 1426.
2. Islam, M.R. and Chakma, A.: Comprehensive Physical and Numerical
Modeling of a Horizontal Well, paper SPE 20627 presented at the 1990
SPE Annual Technical Conference and Exhibition, New Orleans, 2326
September.
3. Folefac, A.N. et al.: Effect of Pressure Along Horizontal Wellbore on
Well Performance, paper SPE 23094 presented at the 1991 Offshore
Europe Conference, Aberdeen, 36 September.
132

4. Ozkan, E., Sarica, C., and Haciislamoglu, M.: Effect of Conductivity


on Horizontal Well Pressure Behavior, paper SPE 24683 presented at
the 1992 SPE Annual Technical Conference and Exhibition, Washington, DC, 47 October.
5. Ihara, M. and Shimizu, N.: Effect of Accelerational Pressure Drop in
a Horizontal Wellbore, paper SPE 26519 presented at the 1993 SPE
Annual Technical Conference and Exhibition, Houston, 36 October.
6. Seines, K. et al.: Considering Wellbore Friction Effects in Planning
Horizontal Wells, JPT (October 1993) 994.
7. Landman, M.J.: Analytical Modeling of Selectivity Perforated Horizontal Wells, J. Petroleum Science and Engineering (1994) 10, 179.
8. Sarica, C. et al.: Influence of Wellbore Hydraulics on Pressure Behavior and Productivity of Horizontal Wells, paper SPE 28486 presented at the 1994 SPE Annual Technical Conference and Exhibition,
New Orleans, 2528 September.
9. Novy, R.A.: Pressure Drops in Horizontal Wells: When Can They be
Ignored? SPERE (1995) 29.
10. Schlichting, H.: Boundary-Layer Theory, seventh edition, J. Kestin
(transl.), McGraw-Hill Book Co., New York City (1987) 378407.
11. Wiesner, M.R. and Chellam, S.J.: Mass Transport Considerations for
Pressure-Driven Membrane Processes, American Water Works Assn. J.
(1992) 84, No. 1, 88.
12. Hartnett, J.P.: Mass Transfer Cooling, Handbook of Heat Transfer
Applications, second edition, W.M. Rohsenow, J.P. Hartnett, and E.N.
Ganic (eds.), McGraw-Hill Book Co., New York City (1985) 1-11111.
13. Berman, A.S.: Effects of Porous Boundaries on the Flow of Fluids in
Systems of Various Geometries Proc., Second U.N. Conference on the
Peaceful Uses of Atomic Energy, Geneva, Switzerland 113 September
(1958) 4, 35158.
14. Yuan, S.W. and Finkelstein, A.B.: Laminar Pipe Flow With Injection
and Suction Through a Porous Wall, Trans., ASME (1956) 78, 719.
15. Kinney, R.B.: Fully Developed Frictional and Heat-Transfer Characteristics of Laminar Flow in Porous Tubes, Intl. J. Heat and Mass
Transfer (1968) 11, No. 9, 1393.
16. Olson, R.M. and Eckert, E.R.G.: Experimental Studies of Turbulent
Flow in a Porous Circular Tube With Uniform Fluid Injection Through
the Tube Wall, J. Applied Mechanics (1966) 33, No. 1, 7.
17. Kinney, R.B. and Sparrow, E.M.: Turbulent Flow, Heat Transfer and
Mass Transfer in a Tube With Surface Suction, J. Heat Transfer (1970)
92, No. 1, 117.
18. Merkine, L, Solan, A., and Winograd, Y.: Turbulent Flow in a Tube
with Wall Suction, J. Heat Transfer (1971) 93, No. 2, 242.
19. Weissberg, H.L. and Berman, A.S.: Velocity and Pressure Distribution
in Turbulent Pipe Flow with Uniform Suction, Proc., Heat Transfer and
Fluid Mechanics Inst., U. of California, Los Angeles, 14 (1955), 130.
20. Doshi, M.R., and Gill, W.N.: Turbulent Flow in a Tube With Wall
Suction, J. Heat Transfer (1974) 96, Series C, No. 2, 251.
21. Lombardi, G., Sparrow, E.M., and Eckert, E.R.G.: Experiments on
Heat Transfer to Transpired Turbulent Pipe Flows, Intl. J. Heat and
Mass Transfer (1974) 17, No. 3, 429.
22. Raithby, G.: Laminar Heat Transfer in the Thermal Entrance Region
of Circular Tubes and Two-Dimensional Rectangular Ducts with Wall
Suction and Injection, Intl. J. Heat and Mass Transfer (1971) 14, No.
2, 223.
23. Kloster, J.: Experimental Research on Flow Resistence in Perforated
Pipe, Masters thesis, Norwegian Inst. of Technology, Trondheim,
Norway (1990).
24. Asheim, H., Kolnes, J., and Oudeman, P.A.: Flow Resistence Correlation for Completed Wellbore, J. Petroleum Science and Engineering
(1992) 8, No. 2, 97.
25. Ihara, M. et al.: Flow in Horizontal Wellbores With Influx Through
Porous Walls, paper SPE 28485 presented at the 1994 SPE Annual
Technical Conference and Exhibition, New Orleans, 2528 September.
26. Su, Z. and Gudmundsson, J.S.: Friction Factor of Perforation Roughness in Pipes, paper SPE 26521 presented at the 1993 SPE Annual
Technical Conference and Exhibition, Houston, 36 October.
27. Yuan, H.J., Sarica, C., and Brill, J.P.: Effect of Perforation Density on
Single Phase Liquid Flow Behavior in Horizontal Wells, paper SPE
37109 presented at the 1996 SPE International Conference on Horizontal Well Technology, Calgary, 1820 November.
SPE Journal, June 1998

28. Ouyang, L.-B., Arbabi, S. and Aziz, K.: General Single Phase Wellbore
Flow Model, topical report, Contract No. DE-FG2293BC14862 U.S.
DOE, Washington, DC (February 1997).
29. Ouyang, L.-B. and Aziz, K.: Steady-State Gas Flow in Pipes, J. Petroleum Science and Engineering (1996) 14, No. 1, 137.
30. Gill, P.E., Murray, W., and Wright, W.H.: Practical Optimization,
Academic Press, London (1981).
31. Ouyang, L.-B.: Stratified Flow Model and Interfacial Friction Factor
Correlations, Masters thesis, Stanford U., Stanford, California (1995).
32. Wallis, G. B. Discussion of Ref. 17, J. Heat Transfer (1970) 92, No. 1,
124.
33. Ouyang, L.-B., Arbabi, S., and Aziz, K.: Preliminary Analysis of the
1995 Stanford Horizontal Wellbore Experiments, Productivity and
Injectivity of Horizontal Wells, annual technical report, Contract No.
DE-FG2293BC14862, U.S. DOE, Washington, DC (March 1996)
124.

SI Metric Conversion Factors

cp 3 1.0*
ft 3 3.048*
lbm 3 4.535 924
psi 3 6.894 757
*Conversion factor is exact.

E203
E201
E201
E100

5
5
5
5

Pazsec
m
kg
kPa
SPEJ

Liang-Biao Ouyang is a PhD degree candidate in petroleum


engineering at Stanford U., Dept. of Petroleum Engineering,
Stanford U., Stanford, California, U. S. A, 94305-2220, e-mail:
ouyang@pangea.stanford.edu. He is working on single-phase
and multiphase wellbore flow modeling, numerical simulation,
and natural-gas engineering. He holds an MS degree in petroleum engineering from Stanford U., and BS and ME degrees in
mechanical engineering from the U. of Science and Technol-

SPE Journal, June 1998

ogy of China. Sepehr Arbabi is a research associate in the


petroleum engineering department at Stanford U., Dept. of
Petroleum Engineering, Stanford U., Stanford, California, U. S. A
94305-2220, e-mail: arbabi@pangea.stanford.edu. His research
interests include horizontal wells, phase behavior, and numerical simulation. He previously spent 1 year as a post-doctoral
fellow at the Reservoir Engineering Research Inst. in Palo Alto,
California. He holds BS and MS degrees in mechanical engineering and a PhD degree in chemical engineering from the U.
of Southern California. Khalid Aziz is a professor of petroleum
engineering at Stanford U., Dept. of Petroleum Engineering,
Stanford U., Stanford, California, U. S. A 94305-2220, e-mail:
aziz@pangea.stanford.edu. His research interests include reservoir simulation, multiphase flow, and horizontal-well technology. He has been an Honorary Member since 1996 and a
Distinguished Member since 1983. He has served on various
section and Society-wide committees, including the Editorial
Review Committee when he was Executive Editor of the SPE
Journal during 199597; and he was a Distinguished Lecturer
from 1987 to 1990. He has been awarded the following honors:
Distinguished Achievement Award for Petroleum Engineering
Faculty (1990), Lester C. Uren Award (1988), Reservoir Engineering Award (1987), and the Cedric K. Ferguson Medal (1979).

Ouyang

Arbabi

Aziz

133

You might also like