Вы находитесь на странице: 1из 9

Sustaining Production by Managing

Annular-Pressure Buildup
A.R. Hasan, SPE, U. of Minnesota-Duluth; B. Izgec, SPE, and C.S. Kabir, SPE, Chevron*

Summary
Increased tubinghead temperature with increased rate may induce
pressure increase in the annuli for the trapped fluid. Managing
annular-pressure buildup (APB) for sustaining well deliverability
is particularly crucial in subsea wells, where intervention is complicated. Ordinarily, a multistring casing design accommodates
anomalous pressure rise from the standpoint of well integrity.
However, management of day-to-day operations presents challenges when APB occurs. This study presents mechanistic models
for understanding and mitigating APB during production. By
preserving mass, momentum, and energy in the wellbore, we
developed two approaches involving semisteady-state and transient
formulations. The intrinsic idea is to mimic the physical process
with minimal input parameters to estimate pressure buildup in the
annuli. Our model formulation handles the mechanisms of fluid
expansion and fluid influx/efflux quite rigorously. This approach
appears to be quite sufficient because we account for most of the
cases of APB encountered.
Introduction
Historically, production and reservoir engineers seldom probed the
root causes of APB, perhaps because tubular design with implicit
APB control has been in the domain of drilling engineers. But the
advent of continuous monitoring of pressure and temperature at
the well bottom, tubinghead, and annuli presents the opportunity
for real-time production and reservoir management within the
safe operating limits of the system. Pressures measured at the
tubinghead and bottomhole with the corresponding flow rate are
the most sought after data in production-engineering calculations.
Rate validation in integrated-asset modeling is a case in point. In
contrast, temperature measurements have not found routine usage,
but are gaining increased attention in connection with transientpressure testing (Sui et al. 2008; Duru and Horne 2008; Izgec
et al. 2007; Hasan et al. 2005; Kabir et al. 1996), downhole flow
profiling (Wang et al. 2008; Nath et al. 2007; Johnson et al. 2006;
Ouyang and Belanger 2006), and flow-rate estimation (Izgec et al.
2009; Kabir et al. 2008). This paper shows that both pressure and
temperature responses at the tubinghead and annuli are strongly
related to flow rates and that these measured values can be used
to alert the operator of possible APB. Naturally, clarity in understanding the interrelationship of wellhead temperature with flow
rate and pressure is imperative for sustaining long-term wellhead
deliverability without compromising well integrity.
Fluid production in a typical production string may conceivably impact pressures in production and surface casings. Generally
speaking, only the shallowest casings are cemented from bottom
to the top, whereas the others are cemented at the bottom, and the
annuli contain mostly drilling fluid. Producing fluids in the tubing
string transmit heat to the liquid-filled annuli, thereby triggering
pressure increase. APB is not associated only with fluid production; fluid circulation during drilling may also induce the same, as
reported by Pattillo et al. (2006). Accessible wellheads in typical
land and offshore dry-tree wells allow the operator to monitor
and bleed off all annuli as needed. However, subsea wells present

*Now with Hess Corporation


Copyright 2010 Society of Petroleum Engineers
Original SPE manuscript received for review 23 August 2008. Revised manuscript received
for review 10 August 2009. Paper (SPE 120778) peer approved 11 October 2009.

May 2010 SPE Production & Operations

serious logistical difficulties for managing APB. Predictably, the


stakes are high in high-pressure/high-temperature (HP/HT) wells
and those that are completed in a deepwater environment where
intervention costs are prohibitive. High flow rates simply exacerbate the APB issue because of the associated high energy that the
fluids bring to surface.
Consequences of APB are increased stresses on tubulars, which
in the limit, may cause well failure. Reduced production to alleviate
annulus-fluid temperature and pressure rise is, of course, economically unattractive. Given this reality, multistring casing designs
have been proposed by Halal and Mitchell (1994) and Adams
and MacEachran (1994). Methods for APB mitigation have been
proposed by way of reduced-heat transport with vacuum-insulated
tubing (Azzola et al. 2007; Ellis et al. 2004), induced compressible
fluids with nitrogen-foam spacers (Vargo et al. 2003, Loder et al.
2003), and water-based spacer fluid containing emulsified methyl
methacrylate monomer (Bloys et al. 2007).
Understanding the underlying mechanisms of APB is essential
for developing any mitigation approach. A number of researchers, notably Adams and MacEachran (1994) and Oudeman and
Bacarreza (1995), have identified three mechanisms responsible
for APBfluid expansion, fluid influx or efflux, and tubing buckling. Mechanistic modeling of APB is currently performed with a
well-known commercial package that relies on numerical modeling of the wellbore heat-transfer process. This study presents two
semianalytical approaches involving semisteady-state and transient
formulations. These models offer the potential for ongoing management of annular pressures. In their current form, these tools
are limited to handling annular-pressure increase owing to fluid
expansion. Issues with tubular buckling are not handled, but fluid
influx or efflux is accounted for in discrete steps.
Model Description
The mass of fluid trapped in an enclosed annulus can experience
significant pressure increase owing to thermal expansion when it
receives heat from the producing fluids in the tubing string. We
can write an expression (Oudeman and Bacarreza 1995; Oudeman
and Kerem 2006) describing the three components contributing to
the annular-pressure increase as
p =

l
1
1
T
Va +
V , . . . . . . . . . . . . . . . . . . . (1)
T
T Va
T Vl l

where T is coefficient of isothermal compressibility, l is coefficient of thermal expansion, Vl is the volume of annular liquid, and
Va is the annular volume. The first term implies liquid expansion,
the second term accounts for volume change in the annulus owing
to tubular buckling, and the third term includes liquid influx (Vl
positive) or efflux (Vl negative) in the annulus. Because the first
term, or the liquid expansion, is by far the most dominant in a
sealed annulus, accounting for well more than 80% of pressure
increase in most cases (Oudeman and Kerem 2006), our modeling
approaches center around this term. In the following, we describe
the development of two methods for estimating APB. In both
methods, we adapt the first term in the Oudeman-Bacarreza model
and integrate it over the entire wellbore.
Semisteady-State Approach. In this approach, we treat uid ow
to be occurring at steady state while the consequent heat exchange
with the formation is in transient mode. This treatment implies
that except for well startup or shut-in when the uid ow occurs
at unsteady state, the use of this method is justied.
195

Consider an annulus of unit depth containing fluid of mass M


per unit length, whose heat content is Mcpi. We use Subscript 1 for
the annulus closest to the tubing (production casing), Subscript 2
for the annulus next to it, and so on. Fig. 1 depicts a typical well
sketch with multiple annuli. In the following, the energy balance
is modeled for Annulus 1; this approach has been extended for
multiple annuli, but we do not address that here for brevity.
During production or shut-in following production, fluids in the
tubing are hotter than the surrounding formation, thereby promoting heat to flow from the tubing toward the formation. Portions of
the heat flowing toward the formation accumulate in various elements of the well, including the annuli. As we mentioned earlier,
heat accumulation with consequent rise in temperature of the fluid
trapped in an annulus causes the APB in the vast majority of the
cases; this element constitutes the models basis.
Heat accumulation in the rst annulus = Heat exiting
the tubing uid Heat going out to the formation
from the rst annulus. . . . . . . . . . . . . . . . . . . . . . . . . . . (2)
Heat accumulation in the annulus is the product of its mass, M,
specific heat, cp1, and change in temperature with time, dT1/dt;
that is, Mcp1(dT1/dt). Heat exiting the tubing fluid to the annulus
is given by the temperature difference between the two fluids and
the heat-transfer coefficient:
Qta = (d1o)U1t (Tf T1). . . . . . . . . . . . . . . . . . . . . . . . . . . . . (3)
Heat loss from the annulus to the formation is modeled in two
steps: that of loss from the annulus fluid to the wellbore/formation
interface and from this interface to earth. The major steps of the
solution are presented here, and the details of the derivation are
presented in Appendix A.
We can write heat loss in terms of temperature difference,
T1Tei, and the relaxation parameter, LR1, which is defined in
Appendix A as
Q = LR1 ( T1 Tei ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (4)
Mathematically, Eq. 2 is then written as
Mcp1(dT1/dt) = (d1o)U1i (Tf T1) LR1(T1 Tei). . . . . . . . . . (5)

ProductionTubing casing
fluid
fluid
4

Fluid (Mud)

2
Cement

In Eq. 5, U1t is the coefficient for heat transfer between the tubing
and the first annular fluid and is detailed in Appendix A. Assuming
the tubing-fluid temperature, Tf , and the heat-transfer coefficients
to be invariant with time makes Eq. 5 a first-order linear-differential equation, with the following solution for the annulus-fluid
temperature, T1, as a function of time:
A A

T1 = T10 e Bt ,
B B

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

where
d U T + LR1Tei
A = ti 1e f
, . . . . . . . . . . . . . . . . . . . . . . . . . . (7)
Mc p1

d U + L R1

B = t1 1e
T1o . . . . . . . . . . . . . . . . . . . . . . . . . . . (8)
Mc p1

In Eq. 8, T1o is the initial temperature of the annulus fluid. Using


the same approach, solutions for fluid temperatures in subsequent
annuli are obtained. This model is semisteady-state because we
presuppose steady-state flow of oil in the flow string, but transient
heat flow in the annuli and formation. Note that Eq. 6 contains the
time-dependant heat-transfer term, LR1, as shown in Eq. A-7 (Hasan
and Kabir 2002).
We use the knowledge of annulus-temperature rise, T1 T10,
to calculate APB. Pressure change, dp, for a fluid in a confined
space is related to the fluids thermal compressibility, , and fluid
expansivity, , by the following expression:
dp =

( V /T ) p
dT = ( / ) dT .
( V /p)T

. . . . . . . . . . . . . . . . . . . . . (9)

The procedure for calculating and and, therefore, the pressure


rise is detailed in Appendix C.
Transient Approach. We adapted our (Izgec et al. 2007) transient
wellbore/reservoir simulator to model APB. While the wellboreuid temperature was evaluated by the coupled simulator, radialheat transfer responsible for APB was accounted for by solving
the following second-order diffusivity equation:
2T 1 T
T
, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (10)
+
=
r 2 r r
t
where represents thermal diffusivity (cp/k), which accounts for
transport of heat through various media, such as fluid, cement,
tubing/casing material, and formation. We adapted our coupled
wellbore/reservoir simulator (Hasan et al. 2005; Izgec et al. 2007)
to model the unsteady-state heat transfer for the APB problem at
hand. The analytic solution of Eq. 10 is given by

3
T f

Formation

wc p LR

(T
mc (1 + C )

ei

Tf +

T f
g sin
+

.
mc p (1 + CT ) z
C p
. . . . . . . . . . . . . . . . . . . . . . . (11)
wc p

The steady-state expression of fluid temperature as a function of


depth is given by
T f
z

= gG sin e ( z L ) LR , . . . . . . . . . . . . . . . . . . . . . . . . . . (12)

where
Fig. 1Well sketch showing various concentric annuli in a
typical well construction.
196

= gG sin +

g sin
. . . . . . . . . . . . . . . . . . . . . . . . . . . . (13)
cp
May 2010 SPE Production & Operations

220

220
Bottomhole

Temperature, F

Temperature, F

Tubinghead

180
Annulus-1

140

Annulus-2

100

180
Midpoint

140
100

Wellhead

60

60

50

100

150

200

250

Time, hr
Fig. 2Time-dependent temperature trace in various annuli
at surface.

and

dp
v dv
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (14)
CJ
dz
c p Jgc dz

a=

wc p
mc p (1 + CT )

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (15)

g sin
= aLRTei aLRT f + a gG sin e ( z L ) LR +
.
t
C p

T f

. . . . . . . . . . . . . . . . . . . . . . . (16)
Integrating Eq. 16, we obtain
T f = Tei +

1 e aLR t
1 e ( z L ) LR , . . . . . . . . . . . . . . . . . . . . (17)
LR

where
LR =

2
wc p

rtoU to ke

. . . . . . . . . . . . . . . . . . . . . . . . . . . (18)
ke + rtoU toTD

The final form of the equation to calculate temperature in


the tubing does not account for heat accumulation or increase in
temperature in the annulus. The general form of the heat-transfer
coefficient used in this formulation is given by
r ln ( rto /r1i ) rto ln ( r co / r ci ) r to ln ( r h / r co )
1
1
=
+ to
+
+
.
kt
kcas
kcem
U to hc + hr
. . . . . . . . . . . . . . . . . . . . . . . (19)
The temperature rise in the annulus will slow down the net
heat-transfer rate. Because we have just one equation to calculate
temperature and heat loss to the formation, the impact of change
in annular heat-transfer rate can be approximated by continuously
evaluating both conductive and convective heat-transfer coefficients in the annulus using the following expressions:
1
1
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (20)
=
U a hc + hr
Heat entering from tubing into the annulus is
Q = 2 rtoU a (T f Ta ). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (21)
May 2010 SPE Production & Operations

0.4

0.6

0.8

Radial Distance, ft
Fig. 3Radial-heat transmission at various points in the wellbore.

Eq. 21 is used to update the annulus-heat-transfer coefficient, Ua.


The initial estimate of the fluid temperature is calculated by using
the following expression:
T f = Tei +

Defining a lumped variable a as

0.2

1 e aLR t
1 e ( z L ) LR . . . . . . . . . . . . . . . . . . . . . (22)
LR

Then, annulus temperature is calculated using the methodology


outlined in Appendix B. After calculation of Ua, the general
overall-heat-transfer coefficient, Uto, is evaluated to estimate the
fluid temperature again. Appendix C outlines how pressures are
computed in other annuli.
Example Applications
1. Synthetic Example. Let us illustrate how temperature propagation occurs radially, leading to APB. Consider a 14,000-ft vertical
well producing single-phase oil with multiple casing strings, such
the one shown in Fig. 1. As expected, the uid temperature at the
tubinghead will be considerably higher than the initial condition
of 70F, owing to energy that the uid brings to surface. Consequently, the tubinghead temperature stabilizes at 200F after
approximately 24 hours of production (Fig. 2). The associated heat
transport in the radial direction also causes temperature increase
in different annuli. Annulus-1 represents the uid temperature
in the production casing. After 24 hours of production, the corresponding energy dissipation in the radial direction is shown in
Fig. 3. As the tubinghead-temperature prole suggests, most of the
temperature drop is locally conned; that is, within the production
casing. These computations were performed with our transient
model, described in the preceeding section. Table 1 presents all
the relevant data used to generate Figs. 1 and 2.
2. Field Example 1. High-rate single-phase oil production occurs
from a 12,000-ft well. Table 2 presents the pertinent input parameters to the simulator. First, we modeled this well behavior with
the semisteady-state approach for pressure increase in 2002 and
2005. Pressure buildup was observed in the 7-in. production casing
owing to heating of annular uid by the producing uid in the tubing string. As Fig. 4 shows, the rise in annulus pressure is directly
related to increase in ow rate. With increased rate, the available
energy for heat transfer increases proportionately, leading to the
increased annular pressure. Fig. 5 makes this point amply clear.
Because the uid cannot expand in the annular conned space, its
low compressibility manifests in terms of increased pressure with
increased temperature, resulting in the APB behavior.
Fig. 6 shows the match of semisteady-state model response
with the wellhead temperature in the tubing. Note that the earlytime mismatch is a result of starting production in a cold well
when in reality, the well had been in production for some time.
However, this mismatch did not impede the models ability to
match the annular-pressure rise, as the pressure match testifies.
Note that the continuous rise in the annular pressure is directly
197

TABLE 2INPUT PARAMETERS FOR FIELD EXAMPLE 1

TABLE 1INPUT DATA FOR GENERATING SYNTHETIC


EXAMPLE
q (STB/D)

q (STB/D)

5,300

3,000

L (ft)

12,000

k (md)

50

L (ft)

14,000

k (md)

700

h (ft)

500

(fraction)

h (ft)

(fraction)

0.2

API Gravity

38

API Gravity

25

Tei (F)

Tei (F)

220

222

ke (Btu/hr-ft-F)

ke (Btu/hr-ft-F)

1.5

1.5

cpe (Btu/lbm-F)

0.625
0.012

cpe (Btu/lbm-F)

0.625

gT (F/ft)

gT (F/ft)

0.011

Tubing OD (in.)

4.5

Tubing OD (in.)

3.5

Tubing ID (in.)

3.96
26

Tubing ID (in.)

2.992

kt (Btu/hr-ft-F)

kt (Btu/hr-ft-F)

30

kc (Btu/hr-ft-F)

kc (Btu/hr-ft-F)

26

30

0.1

ka (Btu/hr-ft-F)

ka (Btu/hr-ft-F)

0.1

kf (Btu/hr-ft-F)

kf (Btu/hr-ft-F)

0.34

(hr/ft )

0.34

(hr/ft )

0.038

0.04

1,400
Annulus Pressure, psig

1,200

Annulus Pressure, psig

300
0.15

1,000
800
600
400
200

5,000

10,000

15,000

20,000

1,000
800
600
400
200
0
170

0
0

1,200

25,000

180

190

200

210

220

Tubinghead Temperature, F

Oil Rate, STB/D

related to the rise in wellhead tubing temperature, which is triggered by increasing flow rate. Fig. 6 captures the essence of what
is shown in Fig. 4.
In 2005, some annular liquid was bled off to relieve pressure. As
a consequence, higher producing rate was restored while the annulus
pressure decreased (Fig. 7). This bleedoff volume was not reported,
however. When we used a bleedoff volume of 79 gal, the model was

able to reproduce the annular-pressure decline quite accurately. Fig. 8


displays both this pressure and temperature match.
To model the entire history involving both pre- and postbleedoff periods, we used both models to compare and contrast
the solutions. As Fig. 9 shows, both models capture the essence
of the annulus-temperature response amid data scatter. Similarly,
good-quality APB match is obtained for the production casing

250

Annulus Pressure, psi

1,400
1,200
1,000

200

800
600
150

400

200

Data
Semisteady-state model

100

0
0

400

800
Time, hr

1,200

Fig. 6Matching tubinghead temperature and APB in early time.


198

Annulus Pressure, psig

Fig. 5Increase in annulus pressure is directly related to increased tubinghead temperature.

Tubinghead Temperature, F

Fig. 4Increasing production rate causes increased heat transfer, leading to APB.

1,600

1,200

Annulus
bleedoff

800

400
0

5,000

10,000 15,000
Oil Rate, STB/D

20,000

25,000

Fig. 7Annulus-pressure bleedoff leads to restoration of high


producing rates in 2006.
May 2010 SPE Production & Operations

1,400

210
205

1,200

200

1,000

195

800

600
400
0

190

Temperature data
Pressure data
Semisteady-state
1,000

2,000

185
180

3,000

4,000

5,000

215
Annulus Pressure, psig

215

Tubinghead Temperature, F

210
205
200
195

185

Data
Semisteady-state
Transient

180
0

Producing Time, hr

Fig. 8Matching tubinghead temperature and annular-pressure buildup in 2006.

(Fig. 10). Note that the late-time pressures reflect bleeding of 79


gal of annular fluid.
3. Field Example 2. This well also exhibits a strong correlation
between tubinghead temperature and APB (Fig. 11). Of course,
changes in ow rate precipitate the corresponding wellhead temperature in the tubing string. Table 3 presents the necessary model
input parameters. Fig. 12 displays a satisfactory match of tubinghead temperature and annular pressure. We used the transient model
to obtain this match. Note that constant rates became the input
because only monthly production rates were available, lacking the
granularity needed for rigorous modeling. Signicant production

1,000

2,000

3,000

4,000

5,000

Producing Time, hr
Fig. 9Comparing tubinghead-temperature predictions in 2006.

curtailment occurred to accommodate APB. Despite the data-quality


issue, the transient models ability to capture overall response is
quite encouraging.
Discussion
Increased drilling in deeper waters makes prediction and management of APB imperative. The two analytic models presented in this
work provide methods for estimating APB. Real-time monitoring
of pressure and temperature in various annuli and the use of proposed models makes day-to-day management of APB much more
viable. In fact, a calibrated model provides clues about the maximum allowable rate commensurate with the systems mechanical
190

1,400
1,200
1,000

800

Data
Semisteady-state
Transient

600
400
0

1000

2000

3000

4000

5000

Tubinghead Temperature, F

1,600
Annulus Pressure, psig

190

Producing Time, hr

180
170
160
150
140
200

400

9,500

L (ft)

14,100

k (md)

34

h (ft)

350

(fraction)

0.15

API Gravity

37

Tei (F)

225

ke (Btu/hr-ft-F)

1.5

cpe (Btu/lbm-F)

0.625

gT (F/ft)

0.011

Tubing OD (in.)

4.5

Tubing ID (in.)

3.96

kt (Btu/hr-ft-F)

26

kc (Btu/hr-ft-F)

26

ka (Btu/hr-ft-F)

0.1

kf (Btu/hr-ft-F)

0.34

(hr/ft )

May 2010 SPE Production & Operations

0.04

1,000

1,200

Fig. 11Good correlation between tubinghead temperature and


annular-pressure rise, Field Example 2.
190
1,100
170
Annular Pressure, psi

q (STB/D)

800

Wellhead Pressure, psig

Fig. 10Comparing APB predictions in 2006.


TABLE 3INPUT PARAMETERS FOR FIELD EXAMPLE 2

600

150
900

130
110

700

o Annular pressure
Tubinghead temperature
Model temperature
Model pressure

500
0

500

1,000

1,500

2,000

90
70

Tubinghead Temperature, F

Annulus Pressure, psig

1,600

50
2,500

Time, hr

Fig. 12Matching tubinghead temperature and annular pressure, Field Example 2.


199

integrity. Moreover, the model can guide an engineer on the bleed


volume needed to improve the operable range of production rate;
Fig. 7 illustrates such an example. The field example shows that
both models allow easy application to managing APB en route to
preserving well integrity. In so doing, aspects of flow assurance
and real-time production management are addressed intrinsically.
Of the two approaches presented, the semisteady-state model
is much simpler than its transient counterpart and, therefore, lends
itself for rapid implementation. However, this simplicity is a tradeoff in that it is not designed to handle transient rates associated
with any production scenario. In contrast, the transient model is
more complex, but is capable of properly representing early-time
transient behavior, as indicated by both the synthetic and field
examples. Consequently, the transient model is capable of providing early warnings for anomalous rise in APB and would allow
ample time for mitigation. Perhaps more important, one can assess
the maximum producing rate at the current condition and after a
certain volume of annular fluid has been bled off.
The examples presented in this study show that applications of
either model require real-time wellhead-pressure and -temperature
data in the production-tubing, production-casing, and other annuli.
Additional input data include well-configuration information, fluid
compressibility, and thermal properties. As is always the case, success in performance prediction is linked directly with data of good
precision for proper management of modern wells.
Conclusions
1. Rigorous formulations of APB are presented. The transient
formulation presents a coupled wellbore/reservoir model,
thereby allowing flexibility in both history matching and future
predictions.
2. Wellhead fluid temperature in the tubing string is directly correlated with APB; the model allows establishing the magnitude
of rate reduction en route to alleviating APB.
3. Field examples illustrate applications of the methodology used,
in terms of both diagnosis and mitigation.
Nomenclature
a = parameter dened by Eq. 14, ft/hr (L/t)
cp = heat capacity, Btu/lbm-F (L2/t2T)
CJ = Joule-Thompson coefcient of uid, ft-F-sec2/lbm (LTt2/M)
CT = thermal-storage coefcient, dimensionless
d = diameter, ft (L)
g = gravitational acceleration, ft/sec2(L/T2)
gc = unit conversion factor (= 32.17 lbm-ft/lbf-sec2)
gG = geothermal gradient, F/ft (T/L)
h = convective heat-transfer coefcient, Btu/F-hr (M/Tt3)
k = thermal conductivity of annular uid, Btu/hr-ft-F (ML/Tt3)
L = length of ow string, ft (L)
LR = relaxation length parameter (Eq. 18), 1/ft1 (L1)
LR1 = relaxation length parameter (Eq. A-7), Btu/ft-hr-oF (ML/Tt3)
M = mass of uid per unit depth, lbm/ft (M/L)
p = pressure, psia (M/Lt2)
q = ow rate, STB/D (L3/t)
Q = heat-transfer rate per unit length of wellbore, Btu/hr-ft
(ML/t3)
Qta = heat-transfer rate from tubing to production annulus per
unit length of wellbore, Btu/hr-ft (ML/t3)
r = radius, ft (L)
rwb = wellbore radius, ft (L)
t = producing time, hr (t)
tD = dimensionless producing time (= ket/ecer2wb)
T = uid temperature, F (T)
Tei = undisturbed earth or formation temperature at any depth,
F (T)
TC = heat transmissivity in center cell, Btu/oF-hr
TD = dimensionless temperature (= 2ke(Twb-Tei)/Q)
TE = heat transmissivity in east cell, Btu/F-hr
TW = heat transmissivity in west cell, Btu/F-hr
200

U = overall heat-transfer coefcient for annular uid, Btu/


hr-F-ft2 (M/Tt3)
V = volume, ft3 (L3)
= coefcient of thermal expansion, 1/F (1/T)
T = coefcient of isothermal compressibility, l/psi (LT2/M)
= density, lbm/ft3 (M/L3)
= parameter dened by Eq. 14, F /ft (T/L)
= diffusivity of heat, hr/ft2 (t/L2)
Subscripts
a = annulus
c = convective
e = earth (formation)
f = produced uid
i = inside
i = ith annulus
o = outside
r = radiative
t = tubing
wb = wellbore
1 = rst or production annulus
Acknowledgment
We are grateful to Chevron management for permission to publish
this work. Author Hasan acknowledges the support received from
the American Chemical Societys Petroleum Research Fund in
conducting work related to this project.
References
Adams, A.J. and MacEachran, A. 1994. Impact on Casing Design of Thermal Expansion of Fluids in Confined Annuli. SPE Drill & Compl 9 (3):
210216. SPE-21911-PA. doi: 10.2118/21911-PA.
Azzola, J.H., Tselepidakis, D.P., Patillo, P.D., Richey, J.F., Tinker, S.J.,
Miller, R.A., and Segreto, S.J. 2007. Application of Vacuum-Insulated
Tubing to Mitigate Annular Pressure Buildup. SPE Drill & Compl 22
(1): 4651. SPE-90232-PA. doi: 10.2118/90232-PA.
Bloys, B., Gonzalez, M., Hermes, R., Bland, R., Foley, R., Tijerina, R.,
Davis, J. et al. 2007. Trapped Annular PressureA Spacer Fluid That
Shrinks. Paper SPE 104698 presented at the SPE/IADC Drilling Conference, Amsterdam, 2022 February. doi: 10.2118/104698-MS.
Duru, O. and Horne, R.N. 2008. Modeling Reservoir Temperature Transients and Matching to Permanent Downhole Gauge Data for Reservoir
Parameter Estimation. Paper SPE 115791 presented at the SPE Annual
Technical Conference and Exhibition, Denver, 2124 September. doi:
10.2118/115791-MS.
Ellis, R.C., Fritchie, D.G. Jr., Gibson, D.H., Gosch, S.W., and Pattillo, P.D.
2004. Marlin Failure Analysis and Redesign: Part 2Redesign. SPE
Drill & Compl 19 (2): 112119. SPE-88838-PA. doi: 10.2118/88838-PA.
Halal, A.S. and Mitchell, R.F. 1994. Casing Design for Trapped Annulus
Pressure Buildup. SPE Drill & Compl 9 (2): 107114. SPE-25694-PA.
doi: 10.2118/25694-PA.
Hasan, A.R. and Kabir, C.S. 2002. Fluid Flow and Heat Transfer in Wellbores. Richardson, Texas: Textbook Series, SPE.
Hasan, A.R., Kabir, C.S., and Lin, D. 2005. Analytic Wellbore-Temperature
Model for Transient Gas-Well Testing. SPE Res Eval & Eng 8 (1):
240247. SPE-84288-PA. doi: 10.2118/84288-PA.
Izgec, B., Hasan, A.R., Lin, D., and Kabir, C.S. 2009. Flow-Rate Estimation
From Wellhead-Pressure and -Temperature Data. SPE Prod & Oper.
SPE-115790-PA (in press; posted 05 November 2009).
Izgec, B., Kabir, C.S., Zhu, D., and Hasan, A.R. 2007. Transient Fluid and
Heat Flow Modeling in Coupled Wellbore/Reservoir Systems. SPE Res
Eval & Eng 10 (3): 294301. SPE-102070-PA. doi: 10.2118/102070-PA.
Johnson, D., Sierra, J., and Gualtieri, D. 2006. Successful Flow Profiling
of Gas Wells Using Distributed Temperature Sensing Data. Paper SPE
103097 presented at the SPE Annual Technical Conference, San Antonio, Texas, USA, 2427 September. doi: 10.2118/103097-MS.
Kabir, C.S., Hasan, A.R., Jordan, D.L., and Wang, X. 1996. A Wellbore/Reservoir Simulator for Testing Gas Wells in High-Temperature
Reservoirs. SPE Form Eval 11 (2): 128134. SPE-28402-PA. doi:
10.2118/28402-PA.
May 2010 SPE Production & Operations

Kabir, C.S., Izgec, B., Hasan, A.R., Wang, X., and Lee, J. 2008. Real-Time
Estimation of Total Flow Rate and Flow Profiling in DTS-Instrumented
Wells. Paper IPTC 12343 presented at the International Petroleum Technology Conference, Kuala Lumpur, 35 December. doi: 10.2523/12343-MS.
Loder, T., Evans, J.H., and Griffith, J.E. 2003. Prediction and Effective Prevention
Solution for Annular Pressure Buildup on Subsea Completed WellsCase
Study. Paper SPE 84270 presented at the SPE Annual Technical Conference
and Exhibition, Denver, 58 October. doi: 10.2118/84270-MS.
Nath, D.K., Sugianto, R., and Finley, D. 2007. Fiber-Optic DistributedTemperature-Sensing Technology Used for Reservoir Monitoring in
an Indonesian Steamflood. SPE Drill & Compl 22 (2): 149156. SPE97912-PA. doi: 10.2118/97912-PA.
Oudeman, P. and Bacarreza, L.J. 1995. Field trial results of annular pressure
behavior in high pressure/high temperature well. SPE Drill & Compl
10 (2): 8488. SPE-26738-PA. doi: 10.2118/26738-PA.
Oudeman, P. and Kerem, M. 2006. Transient Behavior of Annular Pressure
Buildup in HP/HT Wells. SPE Drill & Compl 21 (4): 234241. SPE88735-PA. doi: 10.2118/88735-PA.
Ouyang, L.-B. and Belanger, D. 2006. Flow Profiling by Distributed Temperature Sensor (DTS) SystemExpectation and Reality. SPE Prod &
Oper 21 (1): 269281. SPE-90541-PA. doi: 10.2118/90541-PA.
Pattillo, P.D., Cocales, B.W., and Morey, S.C. 2006. Analysis of an Annular
Pressure Buildup Failure During Drill Ahead. SPE Drill & Compl 21
(4): 242247. SPE-89775-PA doi: 10.2118/89775-PA.
Sathuvalli, U.B., Payne, M.L., Patillo, P., Rahman, S., and Suryanarayana,
P.V. 2005. Development of a Screening System to Identify Deepwater Wells at Risk for Annular Pressure Build-Up. Paper SPE 92594
presented at the SPE/IADC Drilling Conference, Amsterdam, 2325
February. doi: 10.2118/92594-MS.
Sui, W., Zhu, D., Hill, A.D., and Ehlig-Economides, C.A. 2008. Model for
Transient Temperature and Pressure Behavior in Commingled Vertical
Wells. Paper SPE 115200 presented at the SPE Russian Oil and Gas
Technical Conference and Exhibition, Moscow, 2830 October. doi:
10.2118/115200-MS.
Vargo, R.F. Jr., Payne, M., Faul, R., LeBlanc, J., and Griffith, J.E. 2003.
Practical and Successful Prevention of Annular Pressure Buildup on the
Marlin Project. SPE Drill & Compl 18 (3): 228234. SPE-85113-PA.
doi: 10.2118/85113-PA.
Wang, X., Lee, J., Thigpen, B., Vachon, G., Poland, S., and Norton, D. 2008.
Modeling Flow Profile Using Distributed Temperature Sensor (DTS) System. Paper SPE 111790 presented at the Intelligent Energy Conference and
Exhibition, Amsterdam, 2527 January. doi: 10.2118/111790-MS.

Appendix ASolution for Semisteady-State


Approach
Heat accumulation in the annulus is simply equal to heat exiting
the tubing fluid minus the heat going out to the formation. Annular
heat accumulation causes temperature change with time, dT1/dt.
In other words,
Heat accumulation in the annulus = Mcp1(dT1/dt). . . . . . (A-1)
Heat exiting the tubing fluid to the annulus is given by the temperature difference between the two fluids and the heat-transfer
coefficient,
Qta = (d1o)U1t (Tf T1). . . . . . . . . . . . . . . . . . . . . . . . . . . (A-2)
In Eq. A-2, U1t is the coefficient for heat transfer between the
tubing and the first annular fluid and is discussed later in this
section. Heat loss from the annulus to the formation is modeled
in two steps: that of loss from annulus fluid to the wellbore/formation interface and from this interface to earth. Heat loss from the
Annulus 1 fluid (at a temperature of T1) to the wellbore/formation
interface (at Twb) is
Q = 2 r1oU1e ( T1 Twb ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-3)
Heat loss from wellbore/formation interface to earth will depend
on the temperature difference, TwbTei, and the dimensionless temperature TD (Hasan and Kabir 2002) and is given by
May 2010 SPE Production & Operations

2 ke
(Twb Tei ) , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-4)
TD

where TD is given by

TD = ln e(0.2 t ) + (1.5 0.3719e t ) t D . . . . . . . . . . . . . (A-5)


D

Combining Eqs. A-3 and A-4, we obtain


Q = L R1 ( T1 Tei ) , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-6)
where
r1oU1e ke

LR1 2
. . . . . . . . . . . . . . . . . . . . . . . . . (A-7)
ke + ( r1oU1eTD )
Eq. A-2 is then rewritten as
Mcp1(dT1/dt) = (d1o)U1i (Tf T1) LR1(T1 Tei). . . . . . . . (A-8)
In Eq. A-8, U1e represents the overall coefficient for heat transfer
from the first annular fluid to the wellbore/formation interface.
Because heat transfer is more dominant in the radial direction
than in its vertical counterpart, one can treat the resistances to heat
flow in series. Therefore, for heat transfer from Annulus 1 fluid
to the wellbore/formation interface, these resistances include those
offered by (1) Annulus 1 fluid, (2) any insulation around the tubing,
(3) Casing 1 material, (4) Annulus 2 fluid, (5) Casing 2 material,
and (6) the cement sheath, assuming no other annulus. Therefore,
we estimate U1e by inverting the sum of the resistances offered by
all the elements, which is given by
r ln ( rins /r1 ) rto lnn ( r 1o / r 1i )
1
r
r
+
+ to
= to + to
k1
r2 i ha 2
kins
U 1e r1o h1
. . . . . . (A-9)
r to ln ( r 2 o / r 2 i ) r to ln ( r cemo / r 2 o )
+
+
kc 2
kc 2
Eq. A-9 is a general expression. Not all elements, such as tubing
insulation represented by the second term of the right side, may
be present in a given wellbore. Note that all overall-heat-transfer
coefficients in this work are expressed in terms of the tubing outside diameter. Similarly, U1t in Eq. A-2 (and Eq. 5 in the text) is
the coefficient for heat transfer between the tubing and the first
annular fluid. In terms of annular- and tubing-fluid heat-transfer
coefficients, h1 and ht, and tubing conductivity, kt, U1t is given by
r ln ( rto /rti ) 1
1
r
+ . . . . . . . . . . . . . . . . . . . . . (A-10)
= to + to
r
h
kt
ht
U 1t
1o 1
Appendix BSolution of Transient Formulation
The wellbore is represented by cylindrical grids in the z-direction
to calculate heat and mass flow during production and shut-in. The
heat flow from tubing fluid into the formation can be accounted
for by generating radial grids around each cylindrical element and
solving for the conduction equation.
The radial grids around the wellbore are generated logarithmically (Fig. B-1). Heat conduction in an unsteady-state radial
system may be evaluated by performing an energy balance for a
volume element. Using Fouriers law of heat conduction for volume element i, the final form of the equation is given by
Ti n1+1 Ti n +1
T n +1 Ti n +1
+ 2 ki ri h i +1
ri
ri +1
. . . . . . . . . . . . (B-1)
n +1
n
Ti
2 Ti
= c p r h
t
2 ki 1 ri 1h

201

. ... . . .

the formation, and the top of cement occurs inside the annulus of
the previous casing. When the wellhead is sealed, an isolated volume of liquid is created or trapped. Second, a temperature increase
must take place resulting from either production or drilling operations. When the fluid is heated, it begins to expand and can produce
a substantial increase in pressure, which can be compounded if
more than one annulus is isolated.
Basically, the pressure at a specific depth in a trapped column
of liquid is determined by the average annulus temperature, the
volume of annulus, and the amount of fluid trapped. The following
expression for a change in pressure in a contained annulus can be
written by recalling Eq. 1 of the text:
p =

Fig. B-1Radial grids for heat-transfer and pressure-drop


calculations.

In Eq. B-1, the first term represents the heat flux into a volume
element, the second term implies the heat flux out, and the last term
on the right side suggests the heat-accumulation term.
Defining,
TW =

2 ki 1 ri 1h
, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (B-2)
ri

TE =

2 ki ri h
, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (B-3)
ri +1

= c p r 2 h , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (B-4)
Ti n +1 Ti n
. . . . . . . . (B-5)
t
Upon manipulation, the matrix form is given by

TW Ti n1+1 Ti n +1 + TE Ti +n1+1 Ti n +1 =

TW Ti n1+1 TW TE + Ti n +1 TE Ti +n1+1 = Ti n . . . . . . . . (B-6)

t
t
Defining,

TC = TW TE + , . . . . . . . . . . . . . . . . . . . . . . . . . . . . (B-7)

t
n
Ti . . . . . . . . . . . . . . . . . . (B-8)
t
This equation is implicit, which can be solved in matrix form. For
a cylindrical wellbore element with four radial grids, the matrix
from is given by
TW Ti n1+1 TC Ti n +1 TE Ti +n1+1 =

T1n +1

1
TW

TC

TE

TW

TC

TE

TW

TC

T2n +1
T3n +1
T4n +1

Tt n
=

n
T2
. . . . . . . . . . . . (B-9)
t T3n

T4n

The first temperature value on the right side is specified and provided by the analytic temperature solution given in the text.
Appendix CPressure in Multiple Annuli
For a well to experience APB, two conditions are prerequisite.
First, a sealed annulus, or annuli, must exist. In general, a drilled
formation is isolated in a cased well. Cement is circulated above
202

l
1
1
T
V +
V ,
T
T Va a T Vl l

. . . . . . . . . . . . . . . . (C-1)

where is the coefficient of thermal expansion of annular fluid,


is the coefficient of isothermal compressibility, Vl is the volume of
annular liquid, and Va is the annular volume. Eq. C-1 suggests that
three terms contribute to annular-pressure buildup. These are:
Thermal expansion, which results in an increase in pressure
when the annular volume does not increase sufficiently to accommodate this expansion
Change of annular volume, by thermal expansion, ballooning,
or compression of the casings
Change in the amount of the fluid in the annulus caused either
by liquid leakoff to formation or fluid influx into the annulus
In a sealed annulus, the first term, thermal expansion, is dominant. The second term, change in annular volume, is a downward
correction to the first term. For a perfectly sealed annulus, the last
term is eliminated because the amount of liquid remains constant
with time. Because the first term or the liquid expansion is by far
the most dominant in a sealed annulus, accounting for well over
80% of pressure increase in most cases, our modeling approach
centers around this term. For the thermal-expansion term to dominate APB, two conditions need to apply, such that (1) no liquid is
lost from or added to the annulus and (2) the walls of annulus are
completely rigid, a condition that is satisfied when the casings are
cemented to the surface.
Therefore, Eq. C-1 simplifies to the following expression in
absence of leakoff and changes in annular volume:
p =

l
T . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (C-2)
T

The contribution to APB is governed by the ratio of isobaric


thermal expansion and the isothermal compressibility of the fluid.
Thermodynamically, fluid volume (or density) can be expressed as
a function of pressure and temperature, as follows:
V
V
V =
T +
p . . . . . . . . . . . . . . . . . . . . . . . (C-3)
T p
p T
When volume is constant, dV = 0, and we have
V
V
p . . . . . . . . . . . . . . . . . . . . . . . . . . (C-4)

T =
p T
T p
Using the definitions of coefficient of thermal expansion and coefficient of isothermal compressibility, Eq. C-4 can be rearranged
for pressure as

(1/V ) ( V /T ) p (1/ ) ( /T ) p , . . . . . . . . . . . . . . (C-5)


T (1/V ) ( V /p )T (1/ ) ( /p )T , . . . . . . . . . . . . . . . . (C-6)
p =

( V /T ) p
T = ( /T ) T .
( V /p)T

. . . . . . . . . . . . . . . . . . . (C-7)

May 2010 SPE Production & Operations

Note that Eq. C-2 is the finite-difference form of Eq. C-7. Along
with the annular-fluid temperature, fluid properties ( and T) will
change with depth. For example, near the bottomhole, the annular-fluid temperature is high and the temperature increase around
that region with producing time, hence p, is likely to be small.
However, near the wellhead, the annular-fluid temperature is initially low, and may rise substantially over a period of production.
Fig. 3 illustrates this point. Therefore, direct use of Eq. C-7 will
give inaccurate results. To account for the changes in volume of
trapped fluid as a function depth, one needs to use the following
expression:

( V /T ) p
T = ( /T ) T
( V /p)T
.
( M / 2 ) ( /T ) p T {V }bv
=
( M / 2 ) ( /p )T

p =

. . . . . . . . . . . . . . . (C-8)

In Eq. C-8, the summation is taken over the total number of grids
in the simulation model, and M is the mass of trapped fluid for
each grid cell of a given annulus. M can be calculated using the
following expression:
M r 2 /144 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (C-9)
The amount of fluid taken out of the system by bleeding off at any
timestep is represented by the term (V )bv. Note that Eq. C-8 requires
a relevant temperature profile for each annulus at each timestep.

May 2010 SPE Production & Operations

Bulent Izgec is a petroleum engineer at Chevron Energy


Technology Company in Houston. Email: bizg@chevron.com.
His work experience and research areas cover reservoir/production engineering, integrated production modeling, geomechanics, and reservoir simulation. Izgec holds a BS degree in
geophysical engineering from Ankara University, Turkey, and MS
and PhD degrees in petroleum engineering from Texas A&M
University. Rashid Hasan is a professor of chemical engineering at the University of Minnesota-Duluth. He has 30 years of
teaching and research experience in many areas, including
fluid and heat flows in wellbores and pressure-transient testing.
He has consulted with and offered short courses for oil operating and service companies. He has also worked with NASA on
various aspects of multiphase flow and thermohydraulic transients. Hasan has published extensively and has served on various SPE committees, including editorial review for SPEPF and
SPEJ. He holds his MS and PhD degrees from U. of Waterloo,
Canada. Shah Kabir is a senior reservoir engineering advisor
at Hess in Houston. email: skabir@hess.com. His experiences
include transient testing, wellbore fluid- and heat-flow modeling,
and reservoir engineering. Kabir has published two books and
approximately 100 papers. He coauthored the 2002 SPE book
titled Fluid Flow and Heat Transfer in Wellbores. He also contributed in the 2009 SPE monograph on Transient Well Testing. Kabir
holds an MS degree in chemical engineering from the University
of Calgary, Canada. He has served on various SPE committees,
including editorial review committees for SPEPF, SPEREE, and
SPEJ. He received commendation as an outstanding technical
editor five times for two different journals, and also received the
SPE Western Regions Service Award in 2002. He served as SPE
Distinguished Lecturer in 2006-07 and became a Distinguished
Member in 2007.

203

Вам также может понравиться