You are on page 1of 10

Advances in Water Resources 34 (2011) 1656–1665

Contents lists available at SciVerse ScienceDirect

Advances in Water Resources


journal homepage: www.elsevier.com/locate/advwatres

Analytical solution for two-phase flow in a wellbore using the drift-flux model
Lehua Pan a,⇑, Stephen W. Webb b,1, Curtis M. Oldenburg a
a
Lawrence Berkeley National Laboratory, MS-9016, One Cyclotron Road, Berkeley, CA 94720, USA
b
Sandia National Laboratories, P.O. Box 5800, Albuquerque, NM 87185, USA

a r t i c l e i n f o a b s t r a c t

Article history: This paper presents analytical solutions for steady-state, compressible two-phase flow through a well-
Received 25 March 2011 bore under isothermal conditions using the drift flux conceptual model. Although only applicable to
Received in revised form 23 August 2011 highly idealized systems, the analytical solutions are useful for verifying numerical simulation capabili-
Accepted 25 August 2011
ties that can handle much more complicated systems, and can be used in their own right for gaining
Available online 6 September 2011
insight about two-phase flow processes in wells. The analytical solutions are obtained by solving the mix-
ture momentum equation of steady-state, two-phase flow with an assumption that the two phases are
Keywords:
immiscible. These analytical solutions describe the steady-state behavior of two-phase flow in the well-
Wellbore flow
Analytical solution
bore, including profiles of phase saturation, phase velocities, and pressure gradients, as affected by the
Two-phase flow total mass flow rate, phase mass fraction, and drift velocity (i.e., the slip between two phases). Close
Production from wells matching between the analytical solutions and numerical solutions for a hypothetical CO2 leakage prob-
Geologic carbon sequestration lem as well as to field data from a CO2 production well indicates that the analytical solution is capable of
Well leakage capturing the major features of steady-state two-phase flow through an open wellbore, and that the
related assumptions and simplifications are justified for many actual systems. In addition, we demon-
strate the utility of the analytical solution to evaluate how the bottomhole pressure in a well in which
CO2 is leaking upward responds to the mass flow rate of CO2–water mixture.
Ó 2011 Elsevier Ltd. All rights reserved.

1. Introduction numerically mainly because of their nonlinear nature and


complicated relationship between pressure and two-phase velocity
At its most basic level, management of subsurface resources in- fields. On the other hand, analytical solutions, if available, can be
volves a system comprising the wellbore and the target reservoir. used to verify numerical modeling approaches and results. Wallis
As discrete pathways through geologic formations, boreholes and [4] presented a set of analytical solutions for steady-state homoge-
wells are critical to the success of many water, energy, and environ- neous flow (i.e., no slip between the two phases). However, the slip
mental management operations (e.g., geologic carbon sequestra- between the two phases is the fundamental feature that makes two-
tion, oil and gas production, compressed air energy storage, phase flow different from single-phase flow, especially in terms of
geothermal energy production, and subsurface remediation). Simu- flow dynamics. Therefore, analytical solutions restricted to homoge-
lating two-phase flow in wellbores is an important, yet challenging, neous flow conditions provide limited insights into actual two-
task required to answer critical questions on the design and perfor- phase flow systems and have limited applications given the large
mance of fluid production, injection, and transport systems. Because number of two-phase flow situations such as those arising in wells
of the inherent difficulties in modeling two-phase flow with a two- involved with oil and gas, geothermal energy, and geologic carbon
fluid model (e.g., mathematical complications and uncertainties in sequestration, among others.
specifying interfacial interaction terms between the two phases), In this work, we present an analytical procedure to solve the mix-
the drift-flux model, which replaces the separated momentum ture momentum equation of steady-state, compressible, two-phase
equation for each phase by one momentum equation for the mixture flow through a wellbore under isothermal conditions with certain
as a whole complemented with the kinematic constitutive equations simplifications. In addition, we demonstrate that the analytical solu-
specifying the relative motion between phases, was proposed to tions are useful for checking numerical model results for two-phase
significantly reduce these modeling difficulties [1–9]. However, wellbore flow (through an open pipe) and for approximating some
even though the formulation of the drift-flux model is much simpler other practical wellbore processes such as gas leakage up wellbores.
than the two-fluid model, the drift-flux equations are usually solved
2. The drift-flux model and the momentum equation for the
⇑ Corresponding author. Tel.: +1 510 495 2360; fax: +1 510 486 5686. mixture
E-mail address: LPAN@lbl.gov (L. Pan).
1
Current address: Canyon Ridge Consultants LLC, 10 Canyon Ridge Dr., Sandia Park, In the following development, we assume the drift-flux model,
NM 87047, USA. which is limited to one-dimensional flow through an open pipe or

0309-1708/$ - see front matter Ó 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.advwatres.2011.08.009
L. Pan et al. / Advances in Water Resources 34 (2011) 1656–1665 1657

annulus, is appropriate for describing steady-state two-phase flow where Cw is a wall friction factor, d is the wellbore diameter, and Cku
and we develop an analytical solution to solving the drift-flux is a constant.
equations. Variables in the equations below pertaining to quanti- The function K(SG, Ku) in (5) is an interpolation function (from
ties normal to the flow direction should be considered as area- 1.53 to Ku) used to make a smooth transition of drift velocity be-
averaged or assumed to be constant over the cross-section except tween the bubble-rise stage and the film-flooding stage depending
for those explicitly noted otherwise. on gas saturation [12,13], which is similar to the interpolation pro-
The drift-flux models were first developed by Zuber and Findlay posed by Shi et al. (2005) except a nonlinear function is used to en-
[1] and Wallis [4] among others. Although various nomenclatures sure that the first derivative is continue at the switch points.
and forms of equations were used to describe the drift-flux model Note that Eq. (5) described above cannot be applied to the mist
in the literature, the basic idea of the drift-flux models is to assume flow regime, a special two phase flow pattern that often occurs at
that the gas velocity, uG, can be related to the volumetric flux of the high velocity and high gas fraction, X. In the mist flow regime, the
mixture, j, and the drift velocity of gas, ud, by the empirical consti- gas velocity is so high that the small amount of liquid (i.e., X  1)
tutive relationship below: cannot form a film but instead comprises tiny droplets that are uni-
formly distributed in the gas flow. As a result, the slip between the
uG ¼ C 0 j þ ud ð1Þ
two phases in the mist flow regime diminishes. Cheng et al. [14]
where C0 is the profile parameter to account for the effect of local developed a flow pattern map for CO2 that suggests that the mist
gas saturation and velocity profiles over the pipe cross-section. flow regime occurs when mass velocity (i.e., the total mass flow rate
By definition, the volumetric flux j is the volumetrically per unit cross-sectional area, G, in this paper) reaches more than
weighted velocity 300 kg/m2/s at higher X. To account for this region, we suggest to
add to Eq. (5) an adjustment function f(G, X), a smooth function that
j ¼ SG uG þ ð1  SG ÞuL ð2Þ
quickly approaches zero as the state point in the G  X plane gets
where SG is the gas phase saturation. Therefore, the liquid velocity into the mist flow regime whereas it would equal one everywhere
uL can be determined as else (see Appendix B), and calculate the drift velocity as follows:
1  SG C 0 SG ð1  SG Þuc KðSG ; K u ÞmðhÞf ðG; XÞ
uL ¼ j ud ð3Þ ud ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffi ð8Þ
1  SG 1  SG SG qG =qL þ 1  SG
To simplify the problem, we assume that the profile effects over Although the formulations for calculation of the drift velocity
the cross-section are negligible (C0 = 1) as is often the case for are quite complicated, the drift velocity is basically a function of
large-diameter pipe such as typical wellbores [8]. Thus, the stea- Sg, qg, and qL for a given flow system (G and X).
dy-state momentum equation of the mixture in a wellbore with
uniform diameter can be written as follows (see Appendix A):
  3. Analytical solutions
d S qG qL 2 dP Cf qm jum jum
qm u2m þ G ud ¼    qm g cos h
dz 1  SG qm dz 2A At steady state, the total mass flow rate per unit cross-sectional
ð4Þ area G = qmum is a constant and has units of flux, but for conve-
nience we refer throughout to fluxes in the wellbore as flow rates.
Strictly speaking, two-phase flow occurs in different flow re- Therefore, the momentum equation (4) can be rewritten as
gimes resulting in different interfacial interactions. Shi et al. [8] (assuming upward flow for simplicity):
proposed functional forms for the profile parameter and drift !
2
velocity that can be applied continuously for all flow regimes from d G2 SG q G q L 2 dP CfG
þ u ¼   qm g cos h ð9Þ
bubble flow to annular film flow using a set of optimized parame- dz qm 1  SG qm d dz 2Aqm
ters, obtained from an extensive set of large-scale pipe flow
experiments performed by Oddie et al. [10] for one-, two-, and Further, because the solubility of each phase in the other phase
three-phase flows at various inclinations. The following is a sum- is often very low in the systems we are considering, the mass frac-
mary of the mathematical formulations related to the drift velocity tion of the gas phase, X, is also independent of position and the
proposed by Shi et al. [8] (simplified by C0 = 1): densities only depend on pressures and temperature. Therefore,
The drift velocity can be determined as a function of gas satura- by definition, the gas saturation SG (a volumetric fraction) and
tion, the characteristic velocity, uc, the Kutateladze number, Ku (de- the mass fraction X can be related to the total mass flow rate as
fined below), the density of gas phase, qG, the density of liquid follows:
phase qL, and the inclination adjustment, m(h): jG GX=qG X
SG ¼ ¼ ¼ ð10Þ
ð1  SG Þuc KðSG ; K u ÞmðhÞ j þ ud GX=qG þ Gð1  XÞ=qL þ ud X þ a1 qG
ud ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffi ð5Þ
SG qG =qL þ 1  SG where a1 ¼ 1X ud
qL þ G . As a result, the mixture density qm can be writ-
ten as:
The ‘‘characteristic velocity,’’ uc, is a measure of the velocity of
bubble rise in a liquid column, given by X qG a1 qG qL q ðX þ a1 qL Þ
qm ¼ SG qG þ ð1  SG ÞqL ¼ þ ¼ G
 1=4 X þ a1 qG X þ a1 qG X þ a1 qG
g rGL ðqL  qG Þ
uc ¼ ð6Þ ð11Þ
q2L
Or
where rGL is the surface tension between gas and liquid phases and
g is the acceleration of gravity. 1 a1 X 1
¼ þ  ð12Þ
Ku ish the Kutateladze
i number, a function of Bond number, qm X þ a1 qL X þ a1 qL qG
N B ¼ d gðqrL GLqG Þ [11]:
2

From (10), one can obtain the following useful relationship:


" sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi !#12
C ku NB SG X
K u ¼ pffiffiffiffiffiffi 1þ 2 1 ð7Þ ¼ ð13Þ
NB C ku C w 1  SG a1 qG
1658 L. Pan et al. / Advances in Water Resources 34 (2011) 1656–1665

Because qg, and qL are functions of pressure and temperature formation except for a short period at early time. For a system with
here, combining Eqs. (8) and (10) together provides solution of SG high heat capacity fluid (e.g., water), thermal gradients are also of-
and ud for a given pressure and temperature. Similarly, the mixture ten negligible. However, if the dominant fluid is a gas or CO2 the
density, qm, can be obtained using Eq. (11). The derivative on the fluid temperature could decrease with decompression as it flows
left-hand side of Eq. (9), after inserting (13), can be expanded into up the well. As a result, the temperature gradient, although still of-
two parts: ten much smaller than the ambient geothermal gradient, cannot be
      ignored. In this case, we can approximate the temperature profile
d h @ h dP @ h dT
¼ þ ð14Þ along the flow path as piece-wise linear (i.e., dTdz
¼ g T is constant
dz qm @P qm dz @T qm dz within each section) and the entire Eq. (9) can be rewritten as:
where h ¼ G2 þ maG1qL u2d is a function of pressure and temperature,
and T is temperature.      
@ h dP CfG2 @ h
In a well with steady-state upward leakage, the temperature þ1 ¼  qm g cos h  g ð15Þ
@P qm dz 2Aqm @T qm T
gradient tends to be small (a so-called enforced isothermal
condition) because the energy carried by the fluid is often domi- By taking the T and gt as parameters and integrating (15), we
nant over the heat exchange between the well and the surrounding can obtain the following inverse of pressure distribution z(P):

Table 1
Parameters of the two-phase wellbore flow problem.

Parameter Value for analytical solution Value for numerical solution


Length 1000 m (vertical) 1000 m (vertical)
Diameter 0.1 m (circular) 0.1 m (circular)
Total (upward) mass flow rate 50 kg/m2/s (gas + liquid) air: 0.19625 kg/s; water: 0.19625 kg/s (Each = 25 kg/m2/s with a cross sectional area of 7.85  103 m2)
 
Gas mass fraction S q u
0.5 ¼ G GG G
Temperature 20 °C (isothermal) 20 °C (isothermal)
Wellhead pressure 105 Pa 105 Pa
Wall roughness 2.4  105 m 2.4  105 m

Fig. 1. Depth profiles of pressure, gas saturation, gas-phase velocity, and drift velocity under steady-state, isothermal, two-phase (air/water) flow conditions in a vertical
wellbore showing excellent agreement between analytical and numerical results. R2 is the coefficient of determination.
L. Pan et al. / Advances in Water Resources 34 (2011) 1656–1665 1659

Fig. 2. Effects of the drift velocity on the distribution of pressure, gas saturation, and gas-phase velocity. The base case is the same analytical solution as shown in Fig. 1 where
the drift velocity is calculated using Eq. (5). The others assume that the drift velocity is constant over all depths.

 
Z P @
1 þ @P h
qm
  dP 4. Results and discussion
zðPÞ ¼ z0  ð16Þ
CfG2 h
P0
2Aqm
þ qm g cos h þ @
@T q gT
m
4.1. Verification against numerical wellbore models

where P0 is the pressure at the reference location z0. To verify the analytical solution developed above, we first com-
Although a closed-form function could not be obtained except pare the analytical solution to the results of a numerical wellbore
for some special cases, the integration of (16) can be calculated flow simulator T2Well [12,13] which extends TOUGH2 [15] to
using simple numerical integration because the integrand is a handle wellbore flow. We consider an idealized problem of
smooth function of P. The same solution methods could be used steady-state, isothermal, two-phase (air and water) flow through
for the case of (mean) downward flow with a negative sign of a vertical wellbore of 1000 m length. The details of the problem
the friction term because the directions of the friction force and are described in Table 1.
the gravity force are opposite. The T2Well/EOS3 problem is run as a transient problem with
Note that analytical solutions to the momentum equations of adaptive time steps. The ending simulation time is 7.85  108 s
the commonly used simpler models such as constant slip, no slip (4100 steps), at which the average time-derivative of momentum
(i.e., homogeneous flow), or even single phase flow, are all special is about 2.2  1017 (Pa/m) indicating an effective steady state
cases of (16). was reached.
Finally, a general procedure to obtain the analytical solutions In the analytical solution, the gas and liquid phases are pure air
can be summarized as follows: and water, respectively, with the assumption that there is no mixing
between the phases, i.e., the phases are immiscible, an assumption
Step 1: for a series of P (starting from P0), calculate the corre- reasonably well-approximated by phases such as water and air, or
sponding qG(P, T) and qL(P, T), e.g., from equations of state brine and CO2, and somewhat less so by oil and natural gas. The
or look-up tables. phase densities were calculated using the same methods as used in
Step 2: Solve SG and ud from Eqs. (8) and (10) for given qG and the full TOUGH2/EOS3 numerical solution for pure water and air.
qL. However, in the numerical solution, the gas and liquid phases consist
Step 3: Calculate qm, h(P), and the integrand of (16) at each P. of mixtures of water and air components as controlled by a local
Step 4: Calculate z(P) by integration of (16). solubility model.
Step 5: Calculate um( = G/qm) and other variables (e.g., compo- As shown in Fig. 1, the agreement is very good between the
nents of pressure loss) at each point. numerical solution and the analytical solution. The almost perfect
1660 L. Pan et al. / Advances in Water Resources 34 (2011) 1656–1665

match (the coefficient of determination, R2 > 0.998) between ana- pressure is very close to the base case although the pressure match
lytical solutions and the numerical solutions implies that the ef- in the middle regions of the well is slightly off. This holds true for
fects on the two-phase flow of the mixing modeled in TOUGH2 the gas phase velocity also. However, in term of gas saturation, the
of the components in the phases are negligible, at least in this high difference is at the two ends of the borehole (wellhead or bottom-
gas-saturation system. Furthermore, the assumption of isothermal hole). In general, the higher the drift velocity (ud) is, the lower the
conditions in the analytical solution is justified because the up- gas phase velocity (uG) and gas saturation (SG) are. It is easy to
ward-flowing well will quickly establish a nearly steady-state tem- understand the relationship between the drift velocity and the
perature profile with negligible gradient. gas saturation because Eq. (9) implies that the gas saturation is
In this system, although the mass fraction is constant (X = 0.5) inversely proportional to the drift velocity at a given pressure.
throughout the wellbore, the gas saturation decreases with depth However, it is not that straightforward to understand why the
due to pressure increase even though the gas saturation is high gas-phase velocity would decrease as the drift velocity increases
(>0.82) because of the low density of air over the given pressure because the former is suppose to be linearly and positively propor-
range (Fig. 1). Meanwhile, the drift velocity (of the gas phase rela- tional to the latter at given volumetric flux as shown in Eq. (1). The
tive to the mean volumetric velocity) increases with depth from key to understand the phenomena is that the volumetric flux is not
about 0.2 m/s to 0.6 m/s. However, the gas phase velocity de- the same if the drift velocity is different here. Let us rewrite Eq. (1)
creases with depth by about five times over 1000 m (Fig. 1). as following by expressing the volumetric flux in terms of mass
fraction (X), total mass flux (G) and densities of each phase:

4.2. Evaluating the effects of the drift velocity on flow in the wellbore XG ð1  XÞG
uG ¼ j þ ud ¼ þ þ ud ð17Þ
qG qL
The analytical solution developed here can be used to gain sig-
nificant insight into the effects of various assumptions about drift While the gas-phase velocity is linearly proportional to the drift
velocity and gas saturation. Fig. 2 shows the analytical solutions of velocity, it is inversely related to the densities (mainly gas density)
the same problem described in Section 4.1 but with different which will greatly decrease as pressure increases. As a result, the
assumption for the drift velocity. The base case is the analytical increase in the gas-phase velocity due to the increase of the drift
solution where the drift velocity is calculated using Eq. (8). The velocity is overwhelmingly wiped out by the decrease in the gas-
drift velocity of the base case increases almost linearly with depth phase velocity caused by the increase of pressure (in turn the den-
from about 0.2 m/s to 0.6 m/s. If one assumes a constant drift sities of fluids) at the same depth. This also explains why assuming
velocity equal to the average value (=0.4 m/s), the bottomhole the average value of the drift velocity would have a similar effect

Fig. 3. Effects of the drift velocity on the pressure gradients in a steady-state flow wellbore. The base case indicates the analytical solutions where the drift velocity is
calculated using Eq. (5).
L. Pan et al. / Advances in Water Resources 34 (2011) 1656–1665 1661

on the patterns (i.e., results matching at the two ends) of the pres- mixture velocity, for the same total mass flow rate in the homoge-
sure profile and the gas-phase velocity profile as compared to the neous flow case (Fig. 4).
base case.
In addition, higher drift velocity causes higher bottomhole pres- 4.3. Validation with field CO2 production test data
sure for a fixed wellhead pressure (Fig. 2a). In other words, a larger
pressure gradient is needed to maintain the same mass flow rate if To demonstrate how well the analytical solution developed
the drift velocity is higher (Fig. 3). Therefore, drift velocity (or slip) above could be used to describe wellbore flow in an actual well,
is a very important factor in simulating two phase flow problems in particular a well with vulnerable to potential CO2 leakage, we
and any simplification (e.g., assuming homogeneous flow or con- compared our analytical solution to the field data of Cronshaw
stant drift velocity) should be carefully evaluated because it may and Bolling [16]. Flow in the wellbore was believed to reach stea-
lead to a significant mis-estimation of total mass flow for a given dy-state after half a day during a field production test [16], and the
pressure gradient. pressure and temperature data were measured at that time. Totally
Furthermore, as shown in Fig. 3a, the total pressure gradient in- four flow rates were used in the test. The well dimensions and
creases with depth except for the case of homogeneous flow and other parameters are summarized in Table 2 (converted into SI
such increase is mainly caused by the hydrostatic pressure gradi- units).
ent (Fig. 3b). Again, the drift velocity is positively related to the The results of comparison of our analytical solution against
magnitude of such increase. Interestingly, the total pressure gradi- measured field data for this problem are shown in Fig. 5. The close
ent in the homogeneous flow case decreases with increasing depth match (the coefficient of determination is above 0.98 for all cases)
(Fig. 3a). This is caused by the same trends in pressure gradients between the analytical solutions and the measured data for varied
due to both friction loss to the wall and to acceleration (Fig. 3c flow rates indicates that the analytical solution, although strictly
and d) while the hydrostatic pressure gradient is constant applicable only for an idealized system with many simplifications,
(Fig. 3b) in the homogeneous case. For all other cases with non- can capture the major features of this complex CO2–water upward
zero drift velocity, the increase of hydrostatic pressure gradient flow system with an inclined wellbore, even though the system
with depth is a dominant factor over the others (comparing the passes near the critical point of CO2 during upward flow As the
values in Fig. 3a and b), resulting in increase of total pressure gra- flow rate increases, the wellhead pressure decreases and the pres-
dient with depth. Note that the pressure gradient used to over- sure gradient increases at every depth. Such a trend is mainly
come the friction losses to the wall in the homogeneous flow caused by the higher friction loss due to higher fluid velocity and
case is much larger than those in the cases with non-zero drift compensated by the decreased hydrostatic loss due to lighter fluid
velocity. This is because the no-slip condition between the two mixture at lower pressure (Fig. 6). As shown in Fig. 6, the analytical
phases causes higher liquid-phase velocity, and in turn higher solutions indicate that the friction and acceleration pressure

Fig. 4. Effects of the drift velocity on the liquid velocity and the mixture velocity in a steady-state flow wellbore at the given total mass flow rate.

Table 2
Well dimensions and other parameters of the CO2 production test problem.

Parameter Value Note


Length 914.4 m Measured depth
Diameter 0.088 m Circular (tubing)
Incline angle 26.5°
Total (upward) mass flow rate R1 408.95 Gas + liquid
R2 1210.03 Units: kg/m2/s
R3 1827.46 Corresponding to 2.5, 7.4, 11.2, and 13.7 kg/s per well, respectively
R4 2230.20
Gas mass fraction 0.97 Gas mass fraction is defined as equal to SG qGG uG
Temperature From 15.5 to 43.0 °C Measured (8 points) temperature profile along the well for each flow rate
Bottom hole Pressure R1 8.956 Unit: MPa
R2 8.841
R3 8.709
R4 8.586
1662 L. Pan et al. / Advances in Water Resources 34 (2011) 1656–1665

Table 3
Two-phase leakage problem (CO2 and water).

Parameter Value Note


Length 1000 m Vertical wellbore
Diameter 0.1 m Circular
Gas flow rate 0.25–75 (kg/m2/s) 2500 pairs of gas flow
rate and liquid flow rate
Liquid flow rate 0–2500 (kg/m2/s)
Temperature 40 °C Isothermal
Wellhead pressure 0.1 MPa

pressure from the measured steady-state leakage rates of CO2


and water at the surface. The relevance of this problem stems from
the ongoing development and testing of the technology of geologic
carbon sequestration, wherein CO2 from large industrial and
power-plant emission sources will be captured and injected deep
underground to reduce CO2 emissions to the atmosphere (e.g.,
[17]). Despite the great depth, there is a chance that CO2 could leak
up wells that penetrate the storage region (e.g., [18,12]). The spec-
ifications of the two-phase CO2 leakage test problem are described
in Table 3, and results are shown in Fig. 7. Densities of pure CO2
Fig. 5. Results of analytical solution and measured pressures in a CO2 production
well under different flow rates (see Table 2). Lines are analytical solutions whereas
(including supercritical CO2) and aqueous phases as a function of
the symbols are the measured data (red – R1; blue – R2; green – R3; and cyan – R4). pressure and temperature used in the analytical solution were cal-
(For interpretation of the references to colour in this figure legend, the reader is culated using the same methods as in TOUGH2/ECO2N [19].
referred to the web version of this article.) As shown in Fig. 7, the bottomhole pressure is very sensitive to
the gas flow rate whereas it is sensitive to the liquid flow rate only
gradient varies from very small values for lower flow rate to a high
when the gas flow is in the high range. The bottomhole pressure
values that are comparable to the hydrostatic pressure gradient for
decreases as the gas flow rate increases. This is because with high-
higher flow rate, although overall the hydrostatic pressure loss is
er gas-flow rate, the gas saturation is higher and thus the mixture
still a major contribution. In this CO2 dominant (97%) system, the
density is lower, resulting in less hydrostatic pressure at the bot-
temperature decreases as pressure decreases. This effect keeps
tom. As discussed in Section 4.2 (Fig. 3), in such a vertical stea-
the system close to the liquid–gas phase boundary when it be-
dy-state two-phase flow system, the hydrostatic pressure drop is
comes sub-critical. In terms of pressure loss, a lower temperature
often the dominant component of the overall pressure gradient,
contributes to a higher density of CO2 and in turn a higher hydro-
provided that the liquid phase has much higher density than the
static pressure loss. However, the strong dependence of CO2 den-
gas phase. However, this is not to say that a lower pressure plume
sity on pressure is still dominant so that the overall hydrostatic
would have a higher gas leakage rate through an open wellbore. In-
pressure gradient under higher flow rate is still smaller than that
stead, the result only shows that a lower pressure gradient (thus a
under lower flow rate (Fig. 6).
lower down-hole pressure because of the fixed wellhead pressure)
is capable of sustaining a higher gas flow rate in the two-phase
4.4. Estimating the down hole pressure from two-phase leakage flow flowing well. Therefore, a higher pressure plume could drive more
rate at the surface fluid into the well for a given bottomhole pressure, resulting in
higher total flow rates of leakage. Here, one of the critical factors
In this example, we hypothesize a scenario in which CO2 and is the mass fraction of CO2 in the plume, which greatly affects
water are leaking from a well and we show that the analytical the steady-state ratio of gas to liquid phase flow rates. Similar to
solution can be used to estimate the down-hole (bottomhole) gas lifting, more gas could result in not only more gas flow but also

Fig. 6. Analytical solutions of pressure gradients under various flow rates. (a) Hydrostatic; (b) friction + acceleration. R1 – 2.5 kg/s; R2 – 7.4 kg/s; R3 – 11.2 kg/s; and R4 –
13.7 kg/s per well (diameter = 0.088 m).
L. Pan et al. / Advances in Water Resources 34 (2011) 1656–1665 1663

Fig. 7. Contours of bottomhole pressure (Pa) as a function of the mass flow rates of gas and liquid for a well leaking a two-phase mixture of CO2 and water.

more liquid flow. If the wellhead is at atmospheric pressure and The combined-phase momentum equation for wellbore (or
the reservoir pressures is constant in the range of Fig. 7, the con- duct) flow when the axial stress terms are assumed negligible
tour lines of Fig. 7 provide gas leakage rate as a function of liquid can be written as [20]:
leakage rate, and vice versa. ! !
@ X 1 @ X @P Csw
q Sb ub þ A qb Sb u2b ¼   qm g cos h
@t b b A @z b
@z A
5. Concluding remarks
ðA1Þ
This paper shows that with assumptions of negligible mixing where q is density, S is saturation, u is velocity, P is the pressure, A
between two phases and under isothermal conditions, it is possible is cross sectional area of the wellbore, C, is the perimeter of the
to obtain analytical solutions for steady-state two phase flow cross-section, sw is the wall shear stress, and h is the local angle
through a wellbore using the drift-flux model. The resulting analyt- between wellbore section, and the vertical direction. Subscript b
ical solutions are more broadly applicable than prior solutions that indicates phase and m indicates the mixture whereas t is time
assumed homogeneous flow (no slip between the two phases). As and z is spatial coordinate along the wellbore. Note that to be
demonstrated in this work, these analytical solutions are useful for consistent with the drift-flux model, the area-averaged variable,
verifying numerical models of two-phase flow in wellbores that P, is defined as the pressure of the mixture regardless if it is
can simulate a wide variety of flow processes including non-iso- dispersed or film two phase flow because the uniform drift-flux
thermal flows with multiple feed zones, condensation, boiling, model proposed by Shi et al. [8] is applied to all flow regimes with
and other fluid behaviors not amenable to analytical solutions. the same set of the optimized parameters obtained from fitting to
Although the new solutions also have some limiting assumptions, experimental data.
they can also be used to obtain insight into some fundamental The wall shear stress is the friction force between the fluids and
two-phase flow processes in wellbores. the wellbore wall. This term depends on properties and velocities
of both the gas and the liquid phases as well as their fractions of
Acknowledgments the contact area with the wall. Rigorously determining this term
would involve figuring out the detailed two-phase flow structure
This work was supported in part by the CO2 Capture Project near the wall, a difficult task that is intended to be avoided by
(CCP) of the Joint Industry Program (JIP), by the National Risk using the drift-flux model. We assume that the stress is propor-
Assessment Partnership through the Assistant Secretary for Fossil tional to the square of the mixture velocity:
Energy, Office of Sequestration, Hydrogen, and Clean Coal Fuels, 1
National Energy Technology Laboratory (NETL), and by Lawrence sw ¼ f qm jum jum ðA2Þ
2
Berkeley National Laboratory under Department of Energy Contract
where the Fanning friction coefficient (f) is a function of the Rey-
No. DE-AC02-05CH11231. We thank James E. Houseworth (LBNL)
nolds number (Re) ([21], rewritten as Fanning friction coefficient):
for comments on an earlier draft, and three anonymous reviewers
who provided comments that helped us improve the paper. 16
f ¼ for Re < 2400;
Re
Appendix A. Derivation of momentum equation and
  
1 2e=d 5:02 2e=d 13
All variables in the development below should be considered as pffiffiffi ¼ 4 log  log þ for Re > 2400;
f 3:7 Re 3:7 Re
area-averaged or assumed to be constant over the cross-section of
a wellbore except for those explicitly noted otherwise. ðA3Þ
1664 L. Pan et al. / Advances in Water Resources 34 (2011) 1656–1665

where e is the roughness of the wellbore and the Reynolds number By expanding qm and recognizing the relationship (A4), the
is defined as Re = qmumd/lm where lm is the mixture viscosity and d term in [] of (A10) can be simplified as:
is the wellbore diameter.
Before deriving the momentum conservation equation for the SG qG ð1  SG Þqm C 20 þ qL qm ð1  SG C 0 Þ2  ð1  SG Þq2
m
mixture, let us define the mixture density, qm, and the mixture ¼ SG qG ð1  SG Þqm C 20 þ qL qm ð1  SG C 0 Þ2
velocity (velocity of mass center), um, as follows: h i
 ð1  SG Þ S2G q2G C 20 þ q2L ð1  SG C 0 Þ2 þ 2SG qG C 0 qL ð1  SG C 0 Þ
qm ¼ SG qG þ ð1  SG ÞqL ðA4Þ
¼ ð1  SG ÞSG qG C 20 ðqm  SG qG Þ þ qL ð1  SG C 0 Þ2 ðqm  ð1  SG ÞqL Þ
SG qG uG þ ð1  SG ÞqL uL  2SG qG C 0 ð1  SG ÞqL ð1  SG C 0 Þ
um ¼ ðA5Þ h i
qm
¼ SG qG qL ð1  SG Þ2 C 20 þ ð1  SG C 0 Þ2  2C 0 ð1  SG Þð1  SG C 0 Þ
By inserting (1) and (3) into (A5), we can solve j as a function of um
and ud: ¼ SG qG qL ½ð1  SG ÞC 0  ð1  SG C 0 Þ2 ¼ SG qG qL ðC 0  1Þ2 ðA11Þ

qm SG ðqL  qG Þ Similarly, the 2umud term in (A9) can be simplified as:


j¼ u þ ud ðA6Þ
qm m qm  
SG qG qL qm C 0 SG qG qL qm ð1  SG C 0 Þ
where qm ¼ SG C 0 qG þ ð1  SG C 0 ÞqL is the profile-adjusted average 2um ud 
q 2
m ð1  SG Þqm2
density and will reduce to the mixture density if C0 = 1. Note that
the mixture velocity and the volumetric flux of the mixture would
2um ud SG qG qL qm
¼ ½ð1  SG ÞC 0  ð1  SG C 0 Þ
be equal only if there is no slip between two phases (i.e., C0 = 1.0 ð1  SG Þq2m

and ud = 0.0 or homogeneous flow). 2um ud SG qG qL qm


¼ ðC 0  1Þ ðA12Þ
Similarly, the gas velocity and the liquid velocity can also be ex- ð1  SG Þq2m
pressed in terms of um and ud as follows:
And the u2d term can be simplified as:
qm q
uG ¼ C 0 u þ Lu !
qm m qm d SG qG q2L qL S2G q2G SG qG qL u2d
ðA7Þ u2d þ ¼ ½ðð1  SG ÞqL þ SG qG Þ
ð1  SG C 0 Þqm SG qG q2 ð1  SG Þq2 ð1  SG Þq2
uL ¼ um  ud m m m
ð1  SG Þqm ð1  SG Þqm
SG qG qL qm u2d
¼ ðA13Þ
By inserting the stress term (A2) and the phase velocities (A7) ð1  SG Þq2m
into (A1), we can obtain the momentum equation in terms of the
Putting together (A10) through (A13) into (A9), we obtain:
mixture velocity um and the drift velocity ud:
X SG qG qL qm h i
@ 1 @
qb Sb u2b ¼ qm u2m þ ðC 0  1Þ2 u2m þ 2ðC 0  1Þum ud þ u2d
ðqm um Þ þ A qm u2m þ c ð1  SG Þq2
@t A @z b m

@P Cf qm jum jum SG qG qL qm
¼   qm g cos h ðA8Þ ¼ qm u2m þ ½ðC 0  1Þum þ ud 2 ¼ qm u2m þ c ðA14Þ
@z 2A ð1  SG Þq2
m
SG qG qL qm
where the term c ¼ 1S G q2
½ðC 0  1Þum þ ud 2 is caused by slip be- Note that Eq. (A8) is equivalent to the mixture moment equa-
m
tween two phases. tion for the drift model proposed by Hibiki and Ishii (2003) when
While other terms in (A8) are straightforward, the second term the axial stress terms are assumed negligible. When all phases tra-
on the left is obtained as below: vel at the same velocity (i.e., C0 = 1 and ud = 0), c will become zero
X and Eq. (A8) will reduce to the same momentum equation as a sin-
qb Sb u2b ¼ SG qG u2G þ ð1  SG ÞqL u2L
b
gle phase flow. When steady state is reached and C0 = 1, the
 2 momentum Eq. (A8) for flow in a uniform wellbore will be reduced
q qL to an ordinary differential equation (4).
¼ SG q G C 0 m um þ ud
 qm  qm
 2
ð1  SG C 0 Þqm SG q G
þ ð1  SG ÞqL um  ud Appendix B. Adjustment function f(G, X) used in Eq. (8)
ð1  SG Þqm ð1  SG Þqm
!
2 SG qG C 0 qm qL qm ð1  SG C 0 Þ2
2 2 2
The function f(G, X) is a continuous weighting function defined
¼ um þ
q2
m ð1  SG Þq2m on the flow pattern map (G  X) developed by Cheng et al. (2008,
  Fig. 3) in which G is flux (kg m2 s1) and X is gas mass fraction
SG qG qL qm C 0 SG qG qL qm ð1  SG C 0 Þ
þ 2um ud  It shall be zero in the mist flow region where the drift velocity be-
q2
m ð1  SG Þq2
m
! come effectively zero and increases quickly to 1 as the system
2 2 2
S q q
G G L q S q moves away from mist region where the Shi model of drift velocity
þ u2d þ L G G
ðA9Þ
q2
m ð1  SG Þq2 m (Eq. (5) in main text) is optimized.
We simplified the mist flow region boundary on the flow pat-
In (A9), the u2m term can be reorganized as follow: tern map as a straight line which is defined by its two end coordi-
!
SG qG C 20 q2m q q2 ð1  SG C 0 Þ2 nates (Xm1, Gm1), (Xm2, Gm2). A given flow system with a mass flow
2 2
qm um q m um þ u2m þ L m rate G and mass fraction of gas X plots as a point on the flow pat-
q2
m ð1  SG Þq2
m
tern map. Therefore, how far a flow system (G, X) is away from the
qm u2m mist flow region can be measured by the distance from the state
¼ qm u2m þ
ð1  SG Þq2
m point (G, X) to the mist region boundary, (Xm1, Gm1)  (Xm2, Gm2).
h i
According to computational geometry, this distance, Dm (from a
 SG qG ð1  SG Þqm C 20 þ qL qm ð1  SG C 0 Þ2  ð1  SG Þq2
m ðA10Þ
point to a straight line), can be calculated as:
L. Pan et al. / Advances in Water Resources 34 (2011) 1656–1665 1665

Table B1 [4] Wallis GB. One-Dimensional Two-phase Flow. New York: McGraw-Hill Book
Parameters used for adjustment function. Company; 1969.
[5] Hasan AR, Kabir CS. Two-phase flow in vertical and inclined annuli. Int J
Parameter Value Multiphase flow 1992;18(2):279–93.
[6] Ansari AM, et al. A comprehensive mechanistic model for upward two-phase
Xm1 1.0
flow in wellbores. SPEPF (May 1994) 143; Trans. AIME, 1994. p. 297.
Xm2 0.94
[7] Hibiki T, Ishii M. One-dimensional drift-flux model and constitutive equations
Gm1 300 kg/m2/s for relative motion between phases in various two-phase flow regimes. Int J
Gm2 700 kg/m2/s Multiphase flow 2003;46:4935–48.
a 0.001 m2s/kg [8] Shi H, Holmes JA, Durlofsky LJ, Aziz K, Diaz LR, Alkaya B, et al. Drift-flux
k 199 modeling of two-phase flow in wellbores. Soc Pet Eng J 2005:24–33.
[9] Ishii M, Hibiki T. Thermo-Fluid Dynamics of Two-Phase Flow. Springer; 2006.
[10] Oddie G, Shi H, Durlofsky LJ, Aziz K, Pfeffer B, Holmes JA. Experimental study of
X aG 1
two and three phase flows in large diameter inclined pipes. Int J Multiphase

X m1 aGm1 1 Flow 2003;29:527–58.
ðB1Þ [11] Richter HJ. Flooding in tubes and annuli. Int J Multiphase Flow
X m2 aGm2 1 1981;7(6):647–58.
Dm ¼ 1 [12] Pan L, Oldenburg CM, Wu Y-S, Pruess K. Wellbore flow model for carbon
½ðX m2  X m1 Þ2 þ ðaGm2  aGm1 Þ2 2 dioxide and brine. Energy Proc 2009;1(1):71–8. Proceedings of GHGT9,
November 16–20, 2008, Washington DC. LBNL-1416E.
where a is the scaling factor for mass flow rate and jj is the deter- [13] Pan L, Oldenburg CM, Pruess K. T2Well/ECO2N Version 1.0: multiphase and
minant of the matrix. The Dm will become a negative value when non-isothermal model for coupled wellbore-reservoir flow of carbon dioxide
(G, X) is inside of the mist flow region and will increase as the point and water. LBNL-4291E; 2010.
[14] Cheng L, Ribatski G, Quiben JM, Thome JR. New prediction methods for CO2
is ways from the mist flow region. evaporation inside tubes: Part I – A two-phase flow pattern map and a flow
Finally, the adjustment function is defined as: pattern based phenomenological model for two-phase flow frictional pressure
    drops. Int J Heat Mass Transfer 2008;51:111–24.
G [15] Pruess K, Oldenburg CM, Moridis GJ. TOUGH2 user’s guide version 2. E.O.
f ðG; XÞ ¼ max 0:0; 1:0  min 1; expðkDm jDm jÞ ðB2Þ Lawrence Berkeley National Laboratory Report LBNL-43134; 1999.
Gm1
[16] Cronshaw MB, Bolling JD. A numerical model of the non-isothermal flow of
where k is a fitting parameter controlling the steepness of the func- carbon dioxide in wellbores. In: SPE 10735, SPE California regional meeting,
San Francisco, May 22–26, 1982.
tion and the term GGm1 accounts for the fact that the mist flow region [17] IPCC special report on carbon dioxide capture and storage. In: Metz B,
only occurs in the high-velocity region (Cheng et al., 2008, Fig. 3). Davidson O, de Coninck HC, Loos M, Meyer LA, editors. Prepared by Working
The values of the parameters are shown in Table B1. Group III of the Intergovernmental Panel on Climate Change. Cambridge, UK:
Cambridge University Press; 2005.
[18] Gasda SE, Bachu S, Celia MA. Spatial characterization of the location of
References potentially leaky wells penetrating a deep saline aquifer in a mature
sedimentary basin. Environ Geol 2004;46:707–20.
[1] Zuber N, Findlay JA. Average volumetric concentration in two-phase flow [19] Pruess K, Spycher N. ECO2N – a fluid property module for the TOUGH2 code for
systems. J Heat Transfer ASME 1965;87(4):453–68. studies of CO2 storage in saline formations. Energy Convers Manage
[2] Nassos GB, Bankoff SG. Slip velocity ratios in air–water system under steady- 2007;48:1761–7.
state and transient conditions. Chem Eng Sci 1967;22:661. [20] Brennen CE. Fundamentals of Multiphase Flows. Cambridge University Press;
[3] Ishii M. One-dimensional drift-flux model and constitutive equations for 2005.
relative motion between phases in various two-phase flow regimes. Argonne [21] Brill, Mukherjee. Multiphase flow in wells. SPE monograph series, vol. 17.
National Laboratory Report (October 1977) ANL-77-47, 1977. 1999; p. 7, Eq. (2.19).

You might also like