Improved Understanding of Acid Wormholing in Carbonate Reservoirs Through Laboratory Experiments and Field Measurements

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

J191625 DOI: 10.

2118/191625-PA Date: 4-April-20 Stage: Page: 587 Total Pages: 22

Improved Understanding of Acid


Wormholing in Carbonate Reservoirs
through Laboratory Experiments
and Field Measurements
Robert C. Burton, Manabu Nozaki, and Nola R. Zwarich, ConocoPhillips Company,
and Kenji Furui, Waseda University

Summary
A comprehensive study on wormholing has been conducted to improve the understanding of matrix acidizing in carbonate reservoirs.
This work is a continuation of the previous work by Furui et al. (2012a, 2012b). An analysis of additional experimental results, as well
as field measurements, is provided to reinforce and extend the wormhole-penetration model and productivity benefits provided by Furui
et al. (2012b).
A series of small-block tests and one large-block test under geomechanical stresses have been conducted to characterize wormholing
in outcrop-chalk samples. In addition, field data including acid-pumping data and post-stimulation pressure-falloff data have been col-
lected and analyzed to evaluate stimulation effectiveness. Pressure-buildup data from stimulated wells have also been analyzed to eval-
uate the sustainability of the acid-induced skin benefits. Production-logging data have been used to investigate whether created
wormhole networks have remained stable or have collapsed under production stresses. To statistically analyze the data more compre-
hensively, the new data were also compared to field data available in the literature.
The following conclusions are drawn from an analysis of the laboratory data and field data: (1) A skin value of 4 is achievable in
carbonate reservoirs by matrix acidizing; (2) the negative acid skin is relatively stable under production stresses; (3) the wormhole-
penetration model is proved to successfully simulate matrix-acidizing processes in both laboratory-scale and field-scale work; (4) the
small- and large-block laboratory tests reconfirmed wormholing efficiency, which was discussed as a scale effect in the previous stud-
ies; and (5) an understanding of the possible range of wormhole penetration has allowed us to improve field acid treatments and reduce
the risk of connecting to water.
This comprehensive study includes acid-linear-coreflooding tests, small-block tests, large-block tests, and field measurements to
thoroughly analyze acid wormholing in carbonate rock. The database can be very useful information for understanding, benchmarking,
and optimizing future completion/stimulation design.

Introduction
The matrix-acidizing process in carbonate formations is fundamentally different from that in sandstones. Carbonate reservoirs, includ-
ing limestones, chalks, and dolomites, are stimulated with simple acid systems, such as hydrochloric (HCl) acid, while sandstone reser-
voirs are treated with more-complex acid systems containing hydrofluoric (HF) acid or HCl/HF acid mixtures. Historically, it was
believed that carbonate matrix acidizing was a damage-removal process with radial penetration of the acid limited to only a couple of
feet (Daccord et al. 1989; Economides et al. 1994). In some recent papers (e.g., Qiu et al. 2018), this is still claimed. In carbonate
matrix acidizing, the acid does not dissolve the rock uniformly around the wellbore as it is injected into the reservoir. Rather, the acid
finds higher-permeability channels in the pore system and enlarges these channels as it flows outward from the wellbore, forming a net-
work of highly permeable, small passageways called wormholes. Unlike sandstone acidizing, the production performance of carbonate
reservoirs can almost always be improved by acid stimulation treatments even for undamaged wells (e.g., Furui et al. 2012a).
A literature survey carried out for more than 500 wells in Middle Eastern, North Sea, Caspian Sea, Southeast Asian, and North
American fields shows that high-rate matrix-acidizing techniques can be used to effectively stimulate carbonate reservoirs. Field post-
stimulation pressure-buildup and production-test data for carbonate matrix acidizing presented by Furui et al. (2012a) show that skin
factors collected from separate groups of both Middle Eastern limestone and North Sea chalk wells have P50/median skin values in the
3.5 to 4.0 range. Additional public data (Singh 1985; Bartko et al. 1997; Chambers et al. 1997; MaGee et al. 1997; Aslam and Al
Salat 1998; Bazin et al. 1999; Gong and El-Rabaa 1999; Sannier et al. 1999; Al-Dahlan and Nasr-El-Din 2000; Fredd and Miller 2000;
Sharag et al. 2000; Al-Dhafeeri et al. 2002; Thomas and Nasr-El-Din 2003; Bitanov et al. 2005; Albuquerque et al. 2006; Abou-Sayed
et al. 2007; Arangath et al. 2008; Haldar et al. 2008; Rachmawati et al. 2008; Thabet et al. 2009; Foglio and Wtorek 2010; Ussenbayeva
et al. 2012; Van Domelen et al. 2012; Jardim Neto et al. 2013; Clancey et al. 2014; Cui et al. 2014; Folomeev et al. 2014; Kayumov et al.
2014; Issa et al. 2015) and additional ConocoPhillips data have been added to the original data set to provide the plot shown in Fig. 1. As
shown, P50/median skin values from the larger, combined data sets are in the same 3.5 to 4.0 range, as reported previously.
The data set compiled for Fig. 1 covers a wide range of carbonate rock types with reservoir permeability ranging from less than
0.1 md to more than 1,000 md, as shown in Fig. 2.
Review of the information in Figs. 1 and 2 indicates that carbonate matrix-acidizing results are roughly similar over a wide range of
well types, rock types, and reservoir permeability. The poorer skin values in the distribution (skin values worse than 2) are generally
thought to be the result of procedural or operational problems limiting acid placement and/or acid-injection rate. Similarly, the extreme
stimulation results (skin values better than 5) are thought to be the result of acid encountering and then penetrating deeply through
natural-fracture systems in the reservoir. This interpretation of the data implies that well-designed and -executed matrix-acidizing treat-
ments can be expected to provide post-acid skin values in the 2 to 5 range, with P50/median values in the 4 range.

Copyright V
C 2020 Society of Petroleum Engineers

This paper (SPE 191625) was accepted for presentation at the SPE Annual Technical Conference and Exhibition, Dallas, Texas, USA, 24–26 September 2018, and revised for publication.
Original manuscript received for review 18 October 2018. Revised manuscript received for review 1 May 2019. Paper peer approved 30 July 2019.

April 2020 SPE Journal 587

ID: jaganm Time: 14:09 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119


J191625 DOI: 10.2118/191625-PA Date: 4-April-20 Stage: Page: 588 Total Pages: 22

Skin due to Matrix Acid Treatments


654 skin data points from multiple SPE, industry, and ConocoPhillips acid-stimulation references
10

8 All matrix-acidizing skin data from industry and ConocoPhillips references

Skin after Acid Treatment


6
4

–2
–4

–6

–8

–10
0 10 20 30 40 50 60 70 80 90 100
Cumulative % of Population with Skin Worse than Stated Value

Fig. 1—Industry skin data for matrix-acidized carbonate wells.

Skin due to Matrix Acid Treatments


654 skin/permeability data points from multiple SPE, industry, and ConocoPhillips acid-stimulation references
10
8 Reference matrix-acidizing skin value = –4.0
All matrix-acid-treatment skin data from industry and ConocoPhillips references
6
Skin after Acid Treatment

–2
–4

–6

–8

–10
0.01 0.1 1 10 100 1,000 10,000
Reservoir Permeability (md)

Fig. 2—Industry skin data for matrix-acidized carbonate wells vs. reservoir permerability.

Reservoir-normalized acid-treatment volumes used for wells in the data-set range from less than five gal of 28% HCl acid per foot
of formation treated to more than 700 gal of 28% HCl acid per foot of treated interval, with a median acid treatment volume of 75 gal
of 28% HCl/ft and an average volume of 122 gal of 28% HCl/ft. These values are far smaller than would be expected to achieve skin
values in the 4 range on the basis of pore-volume (PV) -based acid injection calculations, implying that only a small fraction of the
total PV is contacted and improved by the acid. These results are in line with the empirical observations of Paccaloni and Tambini
(1993). Acid-injection rates for the data set ranged from less than 0.01 (bbl/min)/ft to more than 1 (bbl/min)/ft, with most jobs in the
0.05- to 0.20-(bbl/min)/ft range.
The productivity benefit of these matrix-acidizing treatments can be estimated using the well-flow-efficiency concept. Well-flow
efficiency (WFE) corresponding to a P50/median skin value of 3.75 for a vertical well with re ¼ 1,056 ft and rw ¼ 0.354 ft is estimated
to be roughly two:
WFE ¼ lnðre =rw Þ=½lnðre =rw Þ þ S ¼ 8:00=½8:00 þ ð3:75Þ ¼ 1:88; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð1Þ

where re is the reservoir-drainage radius in feet, rw is the wellbore radius in feet, and S is the skin factor. This indicates that field matrix
acid treatments tended to improve well productivity from a zero-skin reference value of X (B/D)/psi to a post-acid productivity of
1.88  X (B/D)/psi. Consideration of typical drilling damage providing pre-acid skin values on the order of þ5 to þ 10 indicates that
the true field stimulation benefit is more in the range of

WFE ¼ ½lnðre =rw Þ þ Spre-acid =½lnðre =rw Þ þ Spost-acid  ¼ 3- to 4-fold improvement: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð2Þ

This productivity benefit compares favorably with the stimulation advantages provided by small-to-moderate fracturing treatments.
In a similar manner, Hawkins’ formula (Hawkins 1956) can be used to estimate the effective-acid-stimulation radius from the mea-
sured skin data
Spost-acid ¼ ½k=kpost-acid  1  lnðrwa =rw Þ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð3Þ

where k is the reservoir permeability in md, kpost-acid is the improved permeability because of acid stimulation in md, and rwa is the
acid-stimulation radius or effective wellbore radius. This relation can be simplified to the following equation because the pressure drop
in the acid-stimulation zone is negligibly small:

588 April 2020 SPE Journal

ID: jaganm Time: 14:09 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119


J191625 DOI: 10.2118/191625-PA Date: 4-April-20 Stage: Page: 589 Total Pages: 22

Spost-acid ¼ lnðrwa =rw Þ: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð4Þ

The post-acid skin value can be used to estimate stimulation radius rwa as

rwa ¼ rw  expð1  Spost-acid Þ: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð5Þ

Calculation of rwa values for the acid skin values in Fig. 1 provides the following results: as shown in Fig. 3, the P50/median rwa value
for the data set is approximately 16 ft, indicating significant acid penetration through the formation. This implies that these acid treat-
ments penetrate deeply into the formation, providing a deep stimulation benefit and not just near-wellbore damage remediation.

Skin due to Matrix Acid Treatments


654 acid radius data points from multiple SPE, industry, and ConocoPhillips acid-stimulation references
1,000
All matrix-acidizing skin data from industry and ConocoPhillips references
Effective Wellbore Radius after

100
Treatment (ft)

10

0.1
0 10 20 30 40 50 60 70 80 90 100
Cumulative % of Population with Wellbore Radius Worse than Stated Value

Fig. 3—Stimulation radius calculated from industry skin data for matrix-acidized carbonate wells.

Comparison of the depth of acid stimulation with the small acid volumes used in the treatments implies that only a small fraction of
the total PV is subject to the acid dissolution/wormholing process. Calculations for a median 16-ft stimulation radius interval with a
porosity of 25% show that the total PV is 201 ft3. Comparison of this total PV with the median acid-treatment volume of 75 gal of
28% HCl acid/ft indicates that the acid fills less than 5% of the initial PV and increases the PV by less than 0.8% through the dissolution
of the carbonate-rock matrix.

Literature Review
Laboratory Tests. Numerous experimental studies of wormholing behavior in carbonate acidizing have been conducted, including the
works of Hoefner and Fogler (1988), Wang and Schechter (1993), and Fredd and Fogler (1996). These experimental studies were con-
ducted on the basis of linear core-flow-acidizing tests typically using relatively small cylindrical cores, 1–1.5 in. in diameter and 2.5–6 in.
in length. Results of this experimental work show that there is an optimal injection rate to create the longest wormholes with the minimum
amount of acid. In these studies, the efficiency of matrix acidizing is typically measured in terms of PVs to breakthrough (PVbt) defined as
the pore volumes of the acid needed for wormholes to penetrate or break through the core plug. The smaller the PV to breakthrough, the
smaller the acid volume needed for the wormholes to propagate through a given rock volume and, therefore, the higher the wormhole-
propagation efficiency. Wormholing efficiencies measured for linear core-flow-acidizing experiments using short/small-diameter cores
typically show PVbt values ranging from 1 to 2, with very few instances where PVbt values drop to less than 1.0. It should be noted that
applying these linear core-flow-acidizing test results to field radial-flow treatments on a PV basis results in wormhole lengths/stimulation
radii of only a few feet. These small wormhole-propagation lengths vary significantly from the large acid-stimulation-radius/effective
wellbore values seen from field acid treatments where, as discussed in the preceding section, the majority of rwa values range from 5 to
32 ft with a median value of 16 ft.
It is widely known that wormhole-growth rates, optimal injection rates, and breakthrough volumes depend on the core size used in the
laboratory flow tests. There is a consistent trend showing that wormholing in larger-diameter core samples appears to be more efficient
than that measured in smaller cores, even for relatively homogeneous rock. According to the work presented by Buijse (2000) and Furui
et al. (2012a), it can be found that the wormhole-propagation efficiency is almost proportional to the diameter of the tested core samples.
Intrinsic carbonate heterogeneity, as seen in natural fractures and vugs, is also a very important factor in wormhole propagation. Izgec
et al. (2008) presented linear core-flow-acidizing experiments with 4-in.-diameter vugular limestone cores. The results were compared
with Indiana Limestone data obtained from 1-in.-diameter core samples. Although their results cannot explicitly separate core-size scale
effect and heterogeneity effect, the vugular samples show an order of magnitude lower PVbt values than the relatively homogeneous
Indiana Limestone samples. Their results show that large-scale heterogeneity (vugs) greatly enhances apparent wormholing efficiencies.
To simulate wormhole propagation in a manner closer to field conditions, radial-acid-flow experiments have been performed using
small hollow cylindrical specimens by several investigators (Daccord et al. 1989; Frick et al. 1994; Mostofizadeh and Economides
1994; Tardy et al. 2007; Walle and Papamichos 2015; Karale et al. 2016). With relatively high pump rates, only one dominant worm-
hole was created owing to boundary effects, while with low pump rates, more wormholes were created before one dominant wormhole
broke through. As Gdanski (1999) stated, a symmetrical wormhole structure was observed in the latter, low-pump-rate, cases. McDuff
et al. (2010) performed large-scale acidizing experiments using rock samples up to 14 ft3 in volume and presented computed-
tomography (CT) -scan images of the wormhole network created from a simulated borehole in the large block. The CT images show
that the acid treatment dissolved only a small portion of the carbonate rock to create a radially extending wormhole network around the
borehole. In addition, symmetrical wormhole structures were reconfirmed by their study. According to their work, the acid-injection

April 2020 SPE Journal 589

ID: jaganm Time: 14:09 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119


J191625 DOI: 10.2118/191625-PA Date: 4-April-20 Stage: Page: 590 Total Pages: 22

volume to breakthrough is reported to be 2.5 L, which is equivalent to a PVbt value of 0.03 for an assumed porosity of 20%. This esti-
mated wormhole efficiency matches common field-observed PVbt values of 0.03–0.05. Qiu et al. (2018) performed similar polyaxial-
loading acid-injection tests using 20  16  16-in. Indiana Limestone blocks. They obtained PVbt values of 0.06–0.12.
In summary, large-block testing provides the best match to wormholing efficiencies observed in the field, while the linear core-flow-
acidizing experiments significantly overestimate PVbt values and thus provide very low wormholing efficiencies.

Acid-Wormholing Models. Because of the complex and statistical nature of wormhole propagation, there are several different
approaches taken to model wormhole propagation in carbonate rock. Akanni and Nasr-El-Din (2015) presented a comprehensive
review of the available models. Many models of the wormholing process have been developed including mechanistic models of a
single wormhole or a collection of wormholes (Hung et al. 1989; Schechter 1992), network models (Hoefner and Fogler 1988; Daccord
et al. 1989), and fractal models (Daccord et al. 1989; Pichler et al. 1992). Economides et al. (1994) presented a simple volumetric
model that predicts the wormhole length to which wormholes have propagated around a wellbore as a function of the injected-acid
volume. The model is empirical in nature and requires some input from laboratory experiments or from field experience with a particu-
lar acid system. Buijse and Glasbergen (2005) extended the volumetric model on the basis of the characteristic dependence of the PVs
to breakthrough in acid corefloods on the interstitial velocity. Furui et al. (2012a) augmented the Buijse-Glasbergen model postulating
that the velocity at the tip of the wormholes drives the wormhole-propagation rate and that this interstitial velocity at the tip of the
wormholes is considerably higher than the average interstitial flux calculated from linear core-flow-acidizing experiments. The aug-
mented model matches field observations much better than values predicted by the older models.
Although wormhole structure in linear-flow systems has been studied in detail, there are only a limited number of laboratory studies
on wormhole structure in a radial- or spherical-flow system. Daccord et al. (1989) examined wormhole structures in a radial-flow
system by flowing water through a homogeneous, high-porosity, gypsum material. Frick et al. (1994) and Mostofizadeh and
Economides (1994) flowed acid radially through small (9-cm-diameter), low-permeability limestone samples. In these experiments,
they observed that dominant wormholes develop, while the wormhole patterns are far more complicated and multifaceted in radial-flow
experiments than in the simpler linear core-flow tests. They also pointed out that the interpretation and comparison of radial- and
linear-flow-experiment results require the dimensional transformation of the flow rates between the linear- and radial-flow geometries.
McDuff et al. (2010) observed a similar, radial propagation pattern at an optimal pump rate, while fewer dominant, large wormholes
were found at lower pump rates. Their experimental flow rate corresponds roughly to pumping 0.08 (bbl/min)/ft and generating 10- to
20-ft wormholes into the formation. Abou-Sayed et al. (2007) reached a conclusion similar to that of McDuff, noting that wormhole
lengths of 20–30 ft have been predicted for limestone layers on the basis of core-flow testing, modeling, and well-test results.

Wormhole Stability. The mechanical stability of the wormhole network is crucial to maintaining post-acid well productivity for long
production time periods. Morita et al. (2005) injected 2PV of 15% HCl acid solution into limestone core samples and sliced them for
rock-strength testing using a hardness tester. They concluded that acid reaction reduces the bulk strength of acid-affected zones because
of the porosity loss. However, the remaining grains are still intact and remain tightly bonded, indicating that the local strength should
remain fairly constant, while the borehole inner surface becomes more deformable without shear failure. Walle and Papamichos (2015)
discussed experimental work to investigate wormhole stability using hollow cylindrical samples of Mons chalk. Mons chalk has a cal-
cite content of approximately 99.8% with a very high porosity of 40–44%. They have reported that the wormholes created by acid injec-
tion reduce the rock strength and increase the likelihood of solid production. However, there are no signs of wormhole collapse away
from the near-borehole region.

Current Acid-Wormholing Study. To further improve the understanding of acid wormholing in carbonate reservoirs and optimize
carbonate matrix-acidizing design, laboratory work, field measurements, and design optimization work were performed. The laboratory
work includes (1) linear-acid-coreflooding tests with real-time CT scanning for the characterization of dynamic wormhole propagation,
and (2) large-scale polyaxial-loading acid-flow tests for a validation of the acid-wormholing model described by Furui et al. (2012a,
2012b). The field measurements include (1) post-stimulation skin factors, (2) post-stimulation borehole-size change to confirm deep
acid penetration, and (3) production-logging flow distributions to determine flow changes over time. The design optimization work
involved investigating optimal acid volume, perforation spacing, limited-entry design, and long-term well integrity.

Laboratory Investigations. As discussed in the literature review section, acid linear coreflooding has been used as a standard acid-
wormholing test to characterize acid reaction with carbonate rock, although effects of core dimension on both PVbt and interstitial
velocity (vi) have been recognized. In the current study, the focus has been extended to more-advanced laboratory tests including acid-
wormholing tests under high confining stresses to characterize wormholing more comprehensively.
The following subsections describe each laboratory test briefly and show test results and analysis with the acid-wormholing model
previously presented by Furui et al. (2012a, 2012b).
Linear-Acid-Coreflooding Tests with Real-Time CT Scanning. Previous Linear-Coreflooding Tests. As described, high-porosity
outcrop chalk has been used as an analog core to characterize acid wormholing in Fields A and B. We previously investigated it using
linear-coreflooding tests with different sizes of cores (Furui et al. 2012a) to scale up laboratory linear-flow results to field radial-flow
applications properly. Fig. 4 shows PVbt vs. vi relationships for different acid types (15 wt% HCl acid vs. 28 wt% HCl acid),
different temperatures (150 F vs. 200 F), and different core dimensions (1  6 in. vs. 4  20 in.). The core-test results indicate that a
4-in.-diameter core has PVbt,opt ¼ 0.132, which is a PVbt,opt value approximately four times smaller than that obtained for a 1-in.-diameter
core using 28% HCl acid at 150 F. This partially explains why PVbt is very low in fields. As described by Furui et al. (2012a), P50 PVbt
is typically approximately 0.05 in fields.
Linear-Coreflooding Tests with Real-Time CT Scanning. CT has come to be used as a standard wormhole characterization method
since Bazin et al. (1996) introduced CT to acquire final wormhole structures in a nondestructive manner. Siddiqui et al. (2006) success-
fully characterized wormhole propagation by taking a series of CT images while pumping emulsified acid into a linear core. In this study,
a slightly different experimental setup was used to characterize wormhole propagation using 15% HCl acid without any additives.
Outcrop limestone was used for this test. The core dimension is 4-in. diameter  12-in. length. According to past studies (Bazin
2001; Dong et al. 2014), 6 in. is long enough to obtain a representative result for a core this specific diameter. To obtain a series of CT
images while pumping fluids into a linear core, a core holder was rotated as described in Fig. 5. All laboratory specifications are sum-
marized in Table 1.

590 April 2020 SPE Journal

ID: jaganm Time: 14:09 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119


J191625 DOI: 10.2118/191625-PA Date: 4-April-20 Stage: Page: 591 Total Pages: 22

Linear Acidflooding Test Data for High Porosity Chalk


k = 1–2 md, φ = 30–35%
100
15 wt% HCl at 15°F (1-in. × 6-in.)

Pore Volumes to Breakthrough


28 wt% HCl at 150°F (1-in . ×6-in.)
15 wt% HCl at 200°F (1-in. × 6-in.)
28 wt% HCl at 200°F (1-in. × 6-in.)
10
28 wt% HCl at 150°F (4-in. × 20-in.)

0.1
0.1 1 10
Interstitial Velocity (cm/min)

Fig. 4—Linear-coreflooding-experiment results for high-porosity outcrop-chalk samples (after Furui et al. 2012a).

Rotation direction
Core
Outlet
Core holder

X-ray
detector

Inlet
X-ray
source

Fig. 5—Coreflooding setup with real-time CT scanning.

Core Data
Test 1 Test 2
Rock type Indiana Limestone Indiana Limestone
Unconfined compressive strength ≈5,000 psi ≈5,000 psi
Permeability 31 md 176 md
Porosity 19% 13%
Core diameter 4 in. 4 in.
Core length 12 in. 12 in.
Test Conditions
Test 1 Test 2
Net overburden pressure >300 psi 1,000 psi
Backpressure 800 psi (*) 975 psi
Pump rate 10 mL/min 100 mL/min
Initial saturation fluid 3% KCl 3% KCl
Acid type 15% HCl 15% HCl with 6% Nal
*Unable to maintain 1,000 psi, which is recommended by Cheng et al. (2017) to maintain CO2 in solution.

Table 1—Real-time CT-scanning coreflooding test conditions.

April 2020 SPE Journal 591

ID: jaganm Time: 14:09 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119


J191625 DOI: 10.2118/191625-PA Date: 4-April-20 Stage: Page: 592 Total Pages: 22

15% HCl Acid without Tracer. The first test was conducted as a trial test to investigate whether real-time CT images allow us to
investigate wormhole propagation in real time.
The core sample was saturated with 3% KCl without any tracer. First, untraced water was injected to obtain a stabilized flow at
10 mL/min. Then, water using 1% iodide (as opposed to the 6% NaI planned) tracer was injected while scanning the core in real time.
Once the traced-water breakthrough was confirmed, 15% HCl acid without tracer was injected until acid wormhole breakthrough was
confirmed by the pressure differential becoming negligibly small. Tracer is not typically used to obtain wormhole images.
The 1% iodide tracer concentration was not high enough to capture the injection-fluid front by real-time 2D CT-scan images. The
wormhole propagation was also not captured by real-time 2D CT-scan images. This confirms that effective tracer concentrations would
be required to enhance the CT-number contrast between the wormhole and the matrix. Fig. 6 shows one of the real-time 2D CT-scan
images and a 3D CT image that was obtained by compiling a series of 2D CT images after the coreflooding test.

Outlet

Wormhole

Inlet
2D CT image 3D CT image

Fig. 6—CT images: a real-time CT image obtained while pumping acid (left) and a 3D CT image obtained by compiling a series of
2D images (right) in Test 1.

Although the 2D real-time CT images failed to provide a PVbt, breakthrough was confirmed visually by noting the color change
between the untraced and traced water. The visual breakthrough volume was then used to calculate the water PVbt as 0.53 PVbt at
vi ¼ 1.0 cm/min. This is very close to the acid PVbt of 0.5–0.6 at vi ¼ 1.0 cm/min. This result indicates that there is a preferential,
higher-permeability flow path within the overall PV for this core sample. This also reconfirms that a series of destructive, acid-
coreflooding tests can be replaced by nondestructive laboratory tests such as the tracer tests suggested by Zakaria et al. (2015). Further
study on this subject is encouraged so that field cores can be used for other purposes.
15% HCl acid with 6% NaI. In the second test, a core sample was also saturated with 3% KCl. In the same way as the first test,
3% KCl was pumped for a permeability measurement and the core was scanned by CT to obtain a base image. Then, 3% KCl with
6% NaI was pumped into the core to investigate whether 6% NaI provides CT-number contrasts. Similarly, 3% KCl with 12% NaI was
tried, and, finally, 15% HCl acid with 6% NaI was pumped. To effectively investigate each tracer’s impact on CT numbers, the core
was overflushed by untraced 3% KCl after flowing each fluid.
The 6 and 12% NaI changed CT numbers slightly, although it was difficult to confirm such changes by visual inspection. A preferen-
tial path was not found by the 2D CT images. The permeability in this core was measured to be approximately 5.7 times higher
(170 md vs. 30 md) than that in the previous core. Therefore, it was not possible to obtain water PVbt. Acid PVbt was measured to be
0.3 at vi ¼ 5.1 cm/min.
Use of the 6% NaI improved the CT-number contrast between the core matrix and the wormholes, as shown by Fig. 7. The first four
figures from left are 2D real-time CT images, and the blue lines are wormholes found in the images. The fifth figure is a post-acid 3D
image. As of these tests, only 2D CT images are obtainable with the laboratory configuration, and 2D images cannot fully capture
wormhole propagation. With 2D CT, wormholes that do not propagate in the center of the core might not be visible. Hence, no test was
conducted with the field analog cores that were used in the previous tests (Furui et al. 2012a).
To track wormhole propagation in linear-coreflooding experiments, there have been two alternative methods proposed: (1) pressure
measurements at multiple points along the core (e.g., Qiu et al. 2013) and (2) resistivity measurements at multiple points along the core
(Ghommem et al. 2016). These measurements can show how quickly the wormhole grows. According to Ghommem et al. (2016),
wormhole-propagation rates are not constant for some cases even though acid-injection rate is constant. However, these types of mea-
surements are unable to provide more detail on wormhole-propagation characteristics, such as how the diameter of the
wormhole changes.
Polyaxial-Compression Acidizing Tests. To further understand acid wormholing, larger-scale tests were conducted using cubic
blocks cut from outcrop-chalk samples. These tests focused on (1) the impact of radial/spherical flow on wormhole structure, (2) the
impact of specimen size on PVbt, and (3) wormhole stability under high stresses.
Test Procedure. To simulate matrix acidizing in terms closer to in-situ conditions, cubic chalks were used and polyaxial loads were
applied to the blocks. Two different sizes are used, as described in Fig. 8. For the preperforated liner depicted for the large-block test,
polyvinyl chloride was used. Epoxy was injected as a cement to obtain isolation.

592 April 2020 SPE Journal

ID: jaganm Time: 14:09 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119


J191625 DOI: 10.2118/191625-PA Date: 4-April-20 Stage: Page: 593 Total Pages: 22

Fig. 7—CT images: a real-time CT image obtained while pumping acid (left) and a 3D CT image obtained by compiling a series of
2D images (right) in Test 2.

2-in.

1-in.

Sealing sheet

14-in.
6-in.
14-in.
2-in.
Casing

Cement plug 32-in.


10-in. 3.75-in.
Proppants

Pre-perforated liner
Axial loading plate

27.25-in.

Fig. 8—Block sizes used for polyaxial loading cells. Left: 10 3 10 3 14-in. with a 1-in. hole—small-block test; right: 27.25 3 27.25 3 32-in.
block with a 2-in. hole—large-block test.

The following test steps were taken for each acid-injection test:
1. Drill a hole in the center from the top.
2. Case and cement the hole.
3. Saturate the sample with a saturation fluid.
4. Apply external stresses (r1, r2, and r3 are as defined in Fig. 9), borehole pressure, pore pressure, and temperature.
5. Conduct step-rate test by flowing 3% KCl brine.
6. Inject acid (18% HCl acid) at a specified flow rate or at a specified borehole pressure.
7. Overflush the spent acid with 3% KCl brine.
8. Conduct step-rate test.
9. Increase stresses and pressures.
10. Conduct step-rate test.
11. Release stresses and pressures.
12. Inspect the sample, including CT scanning.
Test Results. Six small-block tests and one large-block test were conducted. The test conditions and results are summarized in
Table 2. Fig. 10 shows the data captured during Test No. 6.

April 2020 SPE Journal 593

ID: jaganm Time: 14:09 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119


J191625 DOI: 10.2118/191625-PA Date: 4-April-20 Stage: Page: 594 Total Pages: 22

σ1 Wellbore

σ2

σ3

σ2 > σ3

Fig. 9—Definitions of external stresses.

Mean Acid
Eff. Acid Pump PVbt
Test Block Sat. Temp. σ1 σ2 σ3 Pressure Stress Rate Permeability Porosity (*)
No. Size Fluid (°F) (psi) (psi) (psi) (psi) (psi) (L/min) (md) (V/V) (V/V)
1 Small OMS 160 4,500 3,000 2,000 1,000 2,167 1.0 N/A N/A 0.051
2 Small Brine 160 2,200 1,200 1,020 1,000 473 1.4 N/A N/A 0.036
3 Small Brine 160 4,500 3,000 2,000 500 2,667 0.4 3.2 N/A 0.055
4 Small Brine 160 3,000 2,500 2,000 1,000 1,500 0.8 3.4 N/A 0.023
5 Small Brine 160 3,000 3,500 2,500 1,000 2,000 0.3–0.9 2 N/A 0.06
6 Small Brine 160 1,250 1,750 1,250 250 1,167 0.4 2.7 N/A 0.036
7 Large Brine 160 2,250 1,500 1,200 1,100 550 0.1–0.5 1.5–2 0.36–0.37 0.024
*When porosity was not measured, porosity is assumed to be 35%.

Table 2—Polyaxial-loading acid-flow-test summary.

10-in. × 10-in. × 14-in. Block Test Flat Jack Volume Measurements


2000 3,000
N–S FJ volume
E–W FJ volume
1500 2,500

Bore and Pore Pressures (psi)


T–B FJ volume
Bore pressure
Flat Jack Volume (cm3)

Pore pressure
1000 2,000

500 1,500

0 1,000

–500 500

–1,000 0
0 50 100 150 200 250 300 350
Elapsed Time (minutes)

Fig. 10—Polyaxial-loading volume measurements for Test No. 6.

Post-Analysis Results. Here, the following four items are discussed using the laboratory test results: (1) the validity of the acid-
wormholing model (Furui et al. 2012b), (2) wormhole structure, (3) wormhole stability, and (4) shear band’s impact on acid wormholing.
Validity of Acid-Wormholing Model. Fig. 11 shows one of the history-matching results. Because the same outcrop rocks as those
tested by Furui et al. (2012a) were used, the wormholing efficiency parameters are assumed to be the same as those used in the previous
work (vi,opt ¼ 1.854 cm/min, PVbt,opt ¼ 0.492, mwh ¼ 6, and c ¼ 1/3). Also, the spherical-flow model was chosen to best fit the test com-
pletion configuration described in Fig. 8. It should also be noted that the model assumes no apparent outer-boundary limit affecting
fluid flow during a test period (i.e., an infinite-acting reservoir). A reasonably good agreement was obtained until a boundary effect
became dominant. Once the dominant wormhole length becomes sufficiently long (i.e., the wormhole approaches very close to the
outer boundary), the flow geometry stops being spherical. Under these conditions, the spherical-flow assumption used in the model is
not valid. In addition, once the wormhole breakthrough has occurred at the outer boundary of the polyaxial block, the direct flow com-
munication from the wellbore to the outer jacket is developed. The current model is not able to simulate these flow behaviors.

594 April 2020 SPE Journal

ID: jaganm Time: 14:09 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119


J191625 DOI: 10.2118/191625-PA Date: 4-April-20 Stage: Page: 595 Total Pages: 22

Test No. 5 Rate History-Matching Result


vi,opt = 1.854 cm/min and PV bt,opt = 0.492
3,000
Bore pressure Rate Rate (model)

2,500

2,000

Rate (mL/min)
Pressure (psi)
1,500

1,000

500

0
295.0 295.1 295.2 295.3 295.4 295.5 295.6 295.7 295.8 295.9 296.0
Elapsed Time (minutes)

Fig. 11—Laboratory rate history-matching example.

Fig. 12 summarizes PVbt values at laboratory and field scale. As a reference data point, limestone test results presented by Tardy
et al. (2007), McDuff et al. (2010), and Qiu et al. (2018) were included (calculated for both 2.5 and 10 L). The field PVbt was calculated
using the following equation:
Vacid
PVbt ¼ 2  r 2 Þh/
; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð6Þ
pðrwh w

where Vacid is the total acid volume pumped in bbl, rw is the wellbore radius, rwh is the wormhole propagation distance, h is the thick-
ness of the reservoir in feet (for horizontal wells, h needs to be changed to L, the interval length of the zone, in feet), and / is the poros-
ity, which is dimensionless. The magnitude of the PVbt in both the small and large laboratory scales is of the same order as field results.
Therefore, these laboratory results are representative of field conditions except for from Tardy et al. (2007) and Qiu et al. (2018). It is
not clear why these results are not in the same range. The specimens used by Tardy et al. (2007) are small compared with the others,
which could have an impact on PVbt just like that seen in the linear-core-flow experiments. In the experiments by Qiu et al. (2018)
(CT images of the test blocks are presented by Aidagulov et al. 2018), multiple wormholes broke through the block, which would not
be expected under laboratory conditions. This multiwormhole breakthrough might explain their high PVbt values.

Comparison of Volumetric Wormholing Efficiency Data


1
Acid Pore Volumes to Breakthrough,
PVbt (V/V)

0.1

0.01
Linear flow Linear flow Radial flow Radial flow Radial Flow Radial flow Field data Linear flow Radial flow
1-in. × 6-in. 4-in. × 20-in. Re = 2.8-in. by 14 ft3 (*) 10-in. × 10-in. 27.25-in. × 1.5 × 12-in. 20-in. × 16-in. ×
chalk chalk H = 2.3-in. Indiana × 14-in. chalk 27.25-in. × Indiana 16-in.
15–28% HCl 28% HCl Indiana Limestone 18% HCl 32-in. chalk Limestone Indiana
150–200°F 150°F Limestone 15% HCl room 18% HCl 15% HCl 98.6°F Limestone
Furui et al. Furui et al. 15% HCl room temperature room Qiu et al. (2018) 15% HCl
(2012a) (2012a) 75–150°F temperature this study temperature 98.6°F
Tardy et al. McDuff et al. this study Qiu et al. (2018)
(2007) (2010)

Fig. 12—Acid pore volumes to breakthrough for the range of laboratory- and field-scale data.

These results also reconfirmed that upscaling is required when linear-coreflooding-test results are used to model field conditions as
examined by several authors (e.g., Buijse 2000).
Wormhole Structure. In this study, CT images were taken for representative blocks using the same methods and CT-scanning equip-
ment as reported by McDuff et al. (2010). Figs. 13a and 13b show CT images of the specimen used in Test 7. The orientation in

April 2020 SPE Journal 595

ID: jaganm Time: 14:09 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119


J191625 DOI: 10.2118/191625-PA Date: 4-April-20 Stage: Page: 596 Total Pages: 22

Fig. 13b is perpendicular to that in Fig. 13a. The wormholes appear to have propagated in a spherical-flow pattern, and one dominant
wormhole propagated more quickly than the other wormholes. As a result, the dominant wormhole broke through to one edge of the
block. This spherical wormhole structure qualitatively validated the spherical-wormhole-propagation model (Furui et al. 2012a). The
dominant wormhole broke through in the maximum-horizontal-stress direction. In other similar tests, the dominant wormhole consis-
tently broke through in the maximum-horizontal-stress, direction even though the specimens were not fractured in these tests. This
might indicate that the stress condition induces some preferential fluid-flow direction.

A dominant
wormhole Spherical
Spherical propagation
propagation pattern
Pre-perforated liner Pre-perforated liner
pattern

(a) CT image in the maximum horizontal stress direction. (b) CT image in the minimum horizontal stress direction.

Fig. 13—CT images obtained on Test 7 block.

Another observation is that wormholes can propagate along the wellbore because of cement-zonal isolation failure. This has been
seen in wells in the field. Some related analyses on field data were performed by Nozaki et al. (2014).
As performed by Schwalbert et al. (2017), numerical analysis might help to provide more insights on how wormholes propagate,
although this type of 2D numerical model has been validated only qualitatively, and further validation work is necessary.
Wormhole Stability. As shown in the CT images in Fig. 13, wormholes were visibly large enough for CT. As concluded by Walle
and Papamichos (2015), the borehole could be unstable in laboratory conditions. By use of a preperforated liner in the large-block test,
borehole-instability issues were not observed. If this borehole instability happens in the field, massive solids production could poten-
tially occur in all wells if rock strength is similar.
Fig. 14 shows the production pressure-drop data from the large-block testing. It shows wormhole stability over the applied stress
range. The skin change was measured to be from 2.6 to 2.5.

Wormhole Stability During Large-Block Testing


Skin = LN(re /rw)*[(acid test DP/zero skin DP) – 1] with re = 13.63-in. and rw = 1.00-in.
1,000
Pre-acid skin = 0
Pressure Drop Across Block During

100
Test (psi)

Pre-acid skin = –2.5


10

Pre-acid skin = –2.6


1

0.1
Pre-acid injection Post-acid injection Post-acid injection
at 50% and 75% stress at 100% stress

Fig. 14—Wormhole stability over applied stress range.

Impact of Shear Bands on Acid Wormholing. In the literature, it was claimed that shear bands improved acid penetration (Gil et al.
2005; Montgomery et al. 2005). To confirm this, a series of permeability measurements was made with different, effective mean
stresses. Fig. 15 shows that step-rate injection-test results show linear relations for q vs. Dp for all tests. No significant injectivity
improvement was observed, although higher effective mean stresses were applied. This indicates that permeability changes caused by
shear bands are minimal.

Field Measurements
Fully Completed Well Design. A set of fully completed, matrix-acid-stimulated wells are reviewed in this section. The wells are com-
pleted by running either an uncemented liner with annulus isolation packers or a cemented liner across the carbonate-reservoir section.

596 April 2020 SPE Journal

ID: jaganm Time: 14:09 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119


J191625 DOI: 10.2118/191625-PA Date: 4-April-20 Stage: Page: 597 Total Pages: 22

Perforations are spaced evenly across the treatment interval, and limited-entry perforation design is used to distribute the acid along the
full reservoir interval. Designed perforation pressure losses ranged from 800 to 1,200 psi for intervals with a uniform reservoir quality
and pressure. For intervals with a high differential pressure across the target reservoir interval, perforation pressure losses are increased
by subdividing the overall treatment length into shorter stage lengths to maximize reservoir-face pressures across the entire zone while
pumping. Long wells are segmented into smaller intervals considering perforation friction, treatment volume, and target acid pump rate
per foot of pay. The acid treatments consist of small volumes of 28% HCl acid injected at a high rate. Typical treatment volumes range
from 75 to 100 gal of 28% HCl acid per foot of completion interval and are pumped at rates in the 0.05- to 0.10-(bbl/min)/ft range. A
typical completion is shown in Fig. 16.

Pre-Acid Injection-Test Results at Different Stress Levels


1,000

900
Measured Pressure Drop (psi)

800

700

600

500 Test 3 pre-acid injection test at J1 = 1,084 psi


Test 3 pre-acid injection test at J1 = 2,167 psi
400
Test 4 pre-acid injection test at J1 = 750 psi
300 Test 4 pre-acid injection test at J1 = 1,500 psi
Test 5 pre-acid injection test at J1 = 1,000 psi
200 Test 5 pre-acid injection test at J1 = 2,000 psi

100 Test 6 pre-acid injection test at J1 = 1,167 psi


Test 7 pre-acid injection test at J1 = 1,150 psi
0
0 50 100 150 200 250 300 350
Pump Rate (mL/min)

Fig. 15—Pre-acid step-rate brine-injection-test results at different stress levels.

Fig. 16—Fully completed well with limited-entry diversion for matrix acid treatment.

Fig. 17 shows a typical acid treatment. The treatment begins with a pre-acid injection test consisting of a step-rate injection test fol-
lowed by a pressure falloff. The step-rate test establishes the injectivity for the treatment interval before the main acid-injection stage
and also provides an estimation of the formation-breakdown pressure. A prestimulation skin factor and formation permeability can also
be estimated from the pre-acid falloff test. If the formation is broken down during the injection test, a fracture closure pressure (i.e.,
minimum horizontal stress) can also be determined from the falloff data.

Typical High-Rate Acid-Treatment Example


9,000 180
Acid on perforations

8,000 160
Fracture closure
pressure

7,000 140
Slurry Rate (bbl/min)

6,000 120
Pressure (psi)
Volume (bbl)

5,000 100
Post-
Step rate test or stimulation
4,000 prestimulation Prestimulation pressure 80
injection test pressure falloff Acid injection Water overflush falloff

3,000 60
Downhole gauge pressure
2,000 40
Reservoir-face pressure

1,000 Total slurry volume 20


Slurry rate
0 0
0 50 100 150
Elapsed Time (minutes)

Fig. 17—A typical, high-rate acid treatment.

April 2020 SPE Journal 597

ID: jaganm Time: 14:09 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119


J191625 DOI: 10.2118/191625-PA Date: 4-April-20 Stage: Page: 598 Total Pages: 22

After conclusion of the pressure-falloff survey, the main acid treatment begins by bullheading the acid to the perforations. This is
performed at a high rate to ensure good flow distribution over the injection interval, generally at greater than fracture pressure. When
the acid hits the perforations, the reservoir-face injection pressure begins to drop as the acid starts to react. Within a few minutes, a
rapid pressure decline occurs. Plotting this information as a function of injection time and reservoir normalized pressure drop allows
skin evolution to be tracked, as shown in the example analysis result in Fig. 18.

Skin-Evolution Plot Example


k = 3.78 md, u = 0.3 cp, Ct = 5.10×10–5 1/psi, and L = 500 ft
6,000
rwa = 0.354
Pressure Drop in the Reservoir (psi) rwa = 1, skin = –1.0
5,000 rwa = 3, skin = –2.1
rwa = 8, skin = –3.1
rwa = 32, skin = –4.5
4,000 rwa = 65, skin = –5.2
First acid at perfs
Acid-treatment data
3,000 First overflush at perfs
Overflush data

2,000

1,000

0
0.001 0.01 0.1 1 10 100
Injection Time (hours)

Fig. 18—Skin-evolution plot during acid-stimulation treatment.

For example, analysis results presented in Fig. 17 above the acid-injection data are plotted in terms of injection time (x-axis) and res-
ervoir pressure loss (y-axis). The pressure drop in the reservoir (i.e., y-axis) is defined by the bottomhole injection pressure at the reser-
voir face minus the initial reservoir pressure before injection. The graph includes differential-pressure type curves calculated from
transient flow equations with various skin factors, while the field injection data, the red and dark blue curves, are plotted on the same
coordinates as the type curves. The results of this comparison can be used for monitoring skin evolution during acid injection.
As shown in the example above, when acid arrives at the reservoir face at 0.35 hours injection time, the effective wellbore radius
is a little less than 8 ft, corresponding to a skin value of 3.1. This negative skin is the result of formation fracturing from high-rate
water injection as the acid is bullheaded to the perforations. As injection time increases to 0.5 hours, the effective wellbore radius has
increased to 15 ft, corresponding to a skin value of 3.8. Skin continues to improve until the end of acid injection at roughly 0.7 hours
when an effective-wellbore-radius value of roughly 32 ft is attained, corresponding to a skin value of 4.5. This final acid skin value
improves as acid is pumped deeper into the formation with the overflush and then holds constant as the majority of the overflush portion
of the treatment is pumped.

Post-Stimulation Borehole Size Measured by Caliper. To investigate acid-penetration depth, a caliper log was run in a horizontal
well’s openhole carbonate interval before and after a coiled-tubing matrix acid treatment. Caliper results are shown in Fig. 19. The orig-
inal borehole diameter was 61/ 8 in. before pumping acid. During the acid-stimulation treatment, 20 gal/ft of 28% HCl acid was pumped
at less than fracture pressure at approximately 2–4 bbl/min through the coiled tubing. Assuming only near-wellbore face dissolution,
with the acid enlarging only the wellbore, the calculated hole size would be 11.84 in. However, the integration of the caliper log shows
the post-acid borehole to have an average diameter of 6.60 in., roughly 0.48-in. diametrical enlargement of the original hole. This small
hole enlargement corresponds to spending only 1.2 gal of 28% HCl acid/ft to enlarge the hole diameter, less than 6% of the 20 gal/ft of
acid pumped. This indicates that even at relatively low-acid injection rates, the majority of the acid moves away from the wellbore to
form wormholes that penetrate farther into the reservoir to provide stimulation.

Differential Caliper Pre- and Post-Acid Stimulation


1.0
Diameter Difference Between

0.8
Calipers (in.)

0.6
Differential caliper
No acid intervals
Average measured
0.4 enlargement diameter

0.2

0
10,500 11,000 11,500 12,000 12,500 13,000 13,500
Measured Depth (ft MDRKB)

Fig. 19—Differential caliper showing borehole enlargement after acid stimulation.

598 April 2020 SPE Journal

ID: jaganm Time: 14:10 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119


J191625 DOI: 10.2118/191625-PA Date: 4-April-20 Stage: Page: 599 Total Pages: 22

Post-Stimulation Skin Evaluation. Skin and acid-contact-length data are the preferred methods to determine the effectiveness of a
carbonate matrix acid treatment. Acid-contact-length data can be readily estimated from perforation friction data collected during treat-
ments pumped using limited-entry fluid-diversion techniques. These acid-contact-length data can also be obtained during or after
the job using production-logging techniques. The depth of acid stimulation and skin can be estimated using job pressure information
(Paccaloni et al. 1988; Prouvost and Economides 1987, 1989; Zhu and Hill 1996) or from post-stimulation pressure falloff (McLeod
and Coulter 1969; Kazemi et al. 1972) and post-production pressure-buildup data. For conventional vertical and inclined wells, reser-
voir k  h and well skin can be readily measured; however, for horizontal wells with longer completion intervals, this can become more
complex. In long horizontal wells, multiple treatments might be performed to achieve stimulation objectives for the entire reservoir-
contact interval, requiring interval-specific transient information to be collected to provide the data needed for effective post-job evalua-
tion. In many cases, this type of data cannot be economically collected; however, the use of intelligent-well systems can readily provide
zone-specific permeability and skin information (Kent et al. 2014). Review of skin values collected during job-pumping operations with
post-stimulation pressure-falloff data and post-production pressure-buildup data collected on wells indicates good agreement in mea-
sured reservoir permeability and well skin data, as shown in Fig. 20.
Post-stimulation skin data for a recent group of North Sea wells designed and executed using the fully completed matrix-acidizing
design process were compared to industry data to determine the benefits of design improvements. This comparison is provided
in Fig. 21.

Skin Calculation Comparison


Skin evolution Post-acid PFO PBU
0
Skin Factor (dimensionless)

–1

–2

–3

–4

–5

–6

Fig. 20—Skin-data comparison among different evaluation methods.

Skin due to Matrix Acid Treatments


645 skin data points from multiple SPE, industry, and ConocoPhillips acid-stimulation references
10

8 All matrix-acidizing skin data from industry and ConocoPhillips references


ConocoPhillips matrix acid treatments following a rigorous design methodology
6
Skin after Acid Treatment

–2

–4

–6

–8

–10
0 10 20 30 40 50 60 70 80 90 100
Cumulative % of Population with Skin Worse than Stated Value

Fig. 21—Comparison of skin data from North Sea matrix-acidized carbonate wells using rigorous design process with general
industry data.

As shown, P50/median skin results for the recent North Sea design process wells are somewhat better than the general industry data,
4.2 vs. 3.7. However, the major benefit of the design process is shown by comparing the acid skin values in the P0–P50 range: the
design process wells have skins ranging from 2.8 to 4.2 with an average of 3.8, while the general industry data trend shows skins
ranging from þ15 to 3.7 with an average of 1.8. The observed skin improvement owing to the implementation of the fully com-
pleted matrix-acidizing design process represents a productivity improvement of 46%,

WFE ¼ ½8:0 þ ðIndustry Savg ¼ 1:8Þ=½8:0 þ ðDesign Process Savg ¼ 3:8Þ ¼ 1:46: . . . . . . . . . . . . . . . . . . . . . . . . . . ð7Þ

April 2020 SPE Journal 599

ID: jaganm Time: 14:10 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119


J191625 DOI: 10.2118/191625-PA Date: 4-April-20 Stage: Page: 600 Total Pages: 22

For a large well program, using the fully completed matrix-acidizing design process yields more-consistent and better skin results
owing to more-effective acid distribution and higher pump rates per foot of pay.

Production-Log Data. To investigate the stability of the wormholes and uniform production distribution along the production interval,
production-logging tools (PLTs) have been run three times in a North Sea carbonate well. This horizontal well was completed using a
fully completed well design and was stimulated in four treatment intervals, each separated by a bridge plug. Table 3 summarizes analy-
sis results of the initial stimulation. Each zone was stimulated with 68–76 gal of 28% HCl acid per foot. By pipe and perforation friction
analysis, all perforations were confirmed to be open; 500–600 psi of perforation friction was achieved at 55–60 bbl/min. Post-
stimulation skin factor was found to range from 4.1 to 4.6, which is slightly better than the P50 skin value observed in Fig. 1.

Friction-Analysis Result Pressure-Transient-Analysis Result

Treatment Acid % of Open Measured


Zone Length Rate Volume Perforations Perforation Permeability Post-Acid Skin
Stage No. (ft-md) (bbl/min) (bbl) (%) Friction (psi) (md) Factor
1 970 55 1,600 100 614 0.5 –4.2
2 990 59 1,600 100 489 1.0 –4.1
3 990 60 1,600 100 614 2.0 –4.2
4 884 60 1,600 100 475 6.0 –4.6

Table 3—Completion, stimulation, and post-stimulation analysis summary.

An initial PLT was run 4 days after the well was put on production. Subsequently, two additional PLTs were run: the first after
2 years of production and the second after 5 years of production. The PLT results are shown in Fig. 22. The three PLTs show that zonal
production contributions have remained steady and in line with job perforation friction analysis and measured reservoir-permeability
data. A majority of the production comes from the Stage 4 interval, which is expected on the basis of the permeability profile along the
lateral measured using prestimulation pressure-falloff data (Table 3). Over the 5 years, there has been no significant change in flow dis-
tribution. Successful acid distribution has been verified through use of on-site perforation friction analysis and by means of post-job
PLT results. The PLT data provide a long-term picture of the effectiveness of the fully completed matrix-acidizing method.

Multiple PLT Results in a Fully Completed Carbonate Well


100
After 4-days production
90
After 16-months production
Production Contribution (%)

80 After 5-years production

70

60

50

40

30

20
10

0
Stage 1 Interval Stage 2 Interval Stage 3 Interval Stage 4 Interval

Fig. 22—Production-logging-tool flow-distribution changeover time in a fully completed well.

A small amount of pressure depletion was confirmed by a series of pressure-buildup data, as shown in Fig. 23. At the time the well
was drilled, the reservoir pressure along the production interval ranged from 3,800 to 4,150 psi as measured using pressure-while-
drilling data. After 50 days of production, 230 days of production, and 818 days of production, the reservoir pressure is estimated to
have dropped to 3,730, 3,630, and 3,400–3,500 psi, respectively. The time-lapse pressure-buildup data suggest changes of k  h/l, as
well as slight changes in skin. There could be multiple possible reasons for the productivity changes to have occurred. On the basis of
the PLT results, the stimulation effectiveness seems to have remained fairly constant, indicating that the wormholes have not collapsed
under the increased net effective stresses applied on them owing to reduction in flowing bottomhole pressures. If the wormholes col-
lapsed in the near-wellbore region, productivity would deteriorate much more significantly. These field observations are consistent with
the wormhole-stability work reported by Morita et al. (2005) and Walle and Papamichos (2015).

Design Considerations
Optimal Acid Volume. As noted in the earlier sections of this paper, operating companies have used a range of acid concentrations
and specific acid volumes (gal/ft) over the years to perform matrix stimulation work in carbonate formations. Over time, specific acid
volumes have declined as operators correlated job skins with acid volumes and moved away from the type of PV calculations associated
with sandstone matrix acidizing.
Williams et al. (1979) recommended 50–200 gal/ft of 28% HCl acid for the matrix treatment of carbonates on the basis of laboratory
work and design studies. Later, Paccaloni and Tambini (1993) reviewed field skin and productivity data from a number of wells and
concluded that 100 gal of 28% HCl acid/ft for carbonate acidizing provided the best balance between cost and productivity gain.

600 April 2020 SPE Journal

ID: jaganm Time: 14:10 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119


J191625 DOI: 10.2118/191625-PA Date: 4-April-20 Stage: Page: 601 Total Pages: 22

Analogous work conducted by ConocoPhillips using well data from several Middle Eastern fields resulted in a similar cost/benefit-
based conclusion, with 28% HCl acid volumes reduced from as high as 1,000 gal/ft to 75 gal/ft over the course of several years. In both
field studies, recommended reductions in acid volume were accompanied by recommendations to increase specific acid injection rate
[(bbl/min)/ft] and provide better, typically mechanical, acid diversion to improve post-acid skin values.

A Series of PBU Data from a Fully Completed Carbonate Well


10,000
Increasing wellbore storage and
wellbore fluid segregation

Rate-Normalized Δp and t*dΔp/dΔt (psi)


and transition to the next flow regime
1,000

100
An indication of
k*h/µ decrease

10

1
Rate-normalized Δp
Rate-normalized t*dΔp/dΔt
0.1 Blue color: After 4-days production
Green color: After 230-days production
Orange color: After 818-days production
0.01
0.00001 0.0001 0.001 0.01 0.1 1 10 100 1,000
Shut-in Time (hours)

Fig. 23—Time-lapse pressure-buildup results from a fully completed carbonate well.

These field studies indicated that pumping large volumes of HCl acid at a low rate provided less stimulation benefit than smaller
acid volumes pumped at higher rates with better diversion. The empirical field-study results were later bolstered by theoretical work in
Hoefner and Fogler (1988) showing the need for high rates to drive acid farther into the formation and improve stimulation.
Current field designs for conventional vertical and inclined wells that focus on deep acid penetration to achieve post-acid skins in
the 4 range tend to use between 75 and 100 gal of 28% HCl acid/ft. For longer horizontal wells, the 75- to 100 gal/ft volumes noted
above are used for jobs requiring deep stimulation to achieve skins in the 4 range, while for simpler acid washes used to achieve skins
in the 2 range, the treatment volume is reduced to 10–20 gal/ft. In all cases, the acid must be mechanically diverted across the entire
reservoir interval using limited-entry perforation designs or packers to achieve effective acid distribution. Further, the acid must be
pumped at a sufficient rate to overcome leakoff and penetrate deeply into the formation.

Impact of Perforation Spacing. The design rate for the fully completed wells discussed in this paper ranges from 0.05 to 0.10 (bbl/min)
of reservoir interval treated. This acid-injection range is based on a combination of leakoff predicted for the reservoir permeability and
the interstitial velocity data obtained from the linear-acid-coreflooding laboratory tests previously discussed. To assess the impact of
increasing the treatment length, thereby reducing the pump rate per foot, a series of comparisons were made using the combined
wormhole-growth model and coupled wellbore/reservoir-flow model described by Furui et al. (2012a, 2012b).
A fully completed horizontal carbonate well was used as the basis for modeling sensitivity calculations. Treatment volumes were
kept consistent at 75 gal of 28% HCl acid and 150 gal of overflush volume per foot of treatment length. Ninety 0.20-in.-diameter perfo-
rations were shot along the treatment length, resulting in a 980-psi perforation pressure loss at the design pump rate of 60 bbl/min.
Treatment length was varied from 1,000 to 4,000 ft, corresponding to perforation spacing ranging from 11 to 45 ft and acid-injection
rates ranging from 0.06 to 0.015 bbl/min of treatment length. Results for a well with uniform reservoir properties along the lateral are
shown in red triangles below in Fig. 24. A permeability of 1 md, a porosity of 0.35, a reservoir pressure of 4,000 psi, and a wellbore
diameter of 8.5 in. were assumed for all cases. Sensitivity runs were performed for a horizontal well with 10 times the permeability in
the heel portion of the lateral compared to the toe, shown by the blue circles. Further sensitivity runs were performed to investigate the
effect of pressure variation along a 1,000-ft lateral, with results shown by the green squares. The modeling results are compared to
field-measured skin values collected from a range of jobs (black points) and for wells acidized with high-perforation-differential-
pressure limited-entry treatments (red points).
The modeling results show that the skin can be improved from 3.2 to 3.6 to 4.0 in a well with uniform reservoir properties
by reducing treatment length from 4,000 to 2,000 to 1,000 ft, corresponding to increasing acid-injection rate from 0.015 to 0.03 to
0.06 (bbl/min)/ft. From a well-flow-efficiency perspective, the skin improvements noted above increase WFE from 1.7 to 1.8 to 2.0. The
various permeability and pressure sensitivity cases similar provide results to those provided by the uniform-reservoir-property cases.
Comparison of the modeling results with field data indicates that increasing acid-injection rate can improve stimulation effectiveness.

Limited-Entry Acid Diversion. The limited-entry stimulation technique was first introduced in 1963 (Lagrone and Rasmusses 1963)
as a method to improve diversion and treat multiple pay intervals in a single treatment. Since that time, numerous papers have been
written about the technique and it has been implemented widely throughout the industry. As mentioned previously, limited-entry perfo-
rating provides the primary means of acid diversion for the fully completed well designs described in this paper with target perforation
pressure losses in the range of 800–1,000 psi. The 800- to 1,000-psi perforation-pressure-loss range has been found to be effective in
allowing breakdown of all perforations where a minimal reservoir-pressure variation along the treatment interval is seen. To validate
this design parameter, a series of field measurements were taken including step-rate tests and running a PLT.
A vertical North Sea chalk well with a cemented liner was stimulated with a high-rate matrix acid stimulation. A total of 58 perfora-
tions with 0.21-in. diameter spaced 7–9 ft apart were shot across three different zones covering a total perforated length of 480 ft. The
deepest zone, Zone C, has three perforations; Zone B has 20 perforations; and the shallowest zone, Zone A, which has the best reservoir

April 2020 SPE Journal 601

ID: jaganm Time: 14:10 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119


J191625 DOI: 10.2118/191625-PA Date: 4-April-20 Stage: Page: 602 Total Pages: 22

properties, has 35 perforations. An initial step-rate test was performed at 60 bbl/min, and the calculated perforation pressure loss indi-
cated that all the perforations were open. During the main treatment, acid was pumped at 60 bbl/min, and good acid-induced pressure
response was observed. The end of job instantaneous shut-in pressure (ISIP) data confirmed all perforations to be open. The post-
stimulation skin factor was calculated to be 4.3, corresponding to an equivalent wellbore radius (rwh) of 21 ft. Before running a PLT
in the well, a step-rate test was performed at injection rates up to 30,000 BWPD. The results of the step-rate ISIP analysis showed that
53 of 58 holes were open (91%) at 30,000 BWPD. The PLT was performed at a 20,000-BWPD surface injection rate, and the results of
the PLT indicated that Zone A was taking 60% of the total injection, Zone B was taking 28%, and Zone C was taking 11%, proportion-
ate to the number of perforations in each of the zones. A comparison between the three injection periods and the flow contribution for
each of the zones is shown in Fig. 25. The injection profile is very consistent between the three injection periods, and the PLT confirms
a uniform distribution across the perforations. This series of field measurements demonstrates that with high acid-treatment rates and
corresponding high perforation pressure losses, entire intervals can be treated effectively.

Field and Modeling Results for Varying Acid-Stimulation-Treatment Lengths


10
Post-Stimulation Skin Factor (dimensionless)

Model results, uniform reservoir properties, 1 md, 4,000 psi reservoir pressure, uncemented
8 Model results, variable permeability,10 (heel)–1 (toe) md, 4,000 psi reservoir pressure, uncemented
Model results, variable reservoir pressure 1 md, 4,000–4,500 (toe) psi reservoir pressure, uncemented
6
Field-measured skin values with variable reservoir properties
4 Field-measured skin values with variable reservoir properties, good limited-entry design

–2

–4

–6

–8

–10
100 1,000 10,000
Stimulated-Interval Length (ft-md)

Fig. 24—Modeling results for single-stage matrix acid stimulation with varying stimulated-interval lengths compared to field data.

Flow Contribution in a Carbonate Well at Different Injection Rates


10,300

10,400
Measured Depth (ft MDRKB)

10,500

10,600

10,700

10,800
Zone A, 315 ft MD interval, 35 perforations, spaced 9 ft apart

10,900 Zone B, 1,454 ft MD interval, 20 perforations, spaced 7 ft apart

Zone C, 20 ft MD interval, 3 perforations, spaced 7 ft apart


11,000
Initial stimulation injection test at 60 bbl/min = 86,400 BWPD surface-injection rate

11,100 PLT at 20,000 BWPD surface-injection rate

Step-rate test at 30,000 BWPD surface-injection rate


11,200
0 10 20 30 40 50 60 70 80 90 100
Injection Flow Contribution (%)

Fig. 25—An example of limited-entry fluid-diversion measurements.

Importance of Acid Diversion for Well Integrity. Casing deformation has a long history and is a reality for many operators in highly
compacting carbonate reservoirs, with the deformations occurring in both the overburden and reservoir formations (Yudovich et al.
1989; Anvik and Gibson 1987; Barkved et al. 2003). Attempts to minimize failure frequency and magnitude through reservoir-pressure
maintenance, increasing drilling angle, and increasing the casing size as Yudovich et al. (1989) have suggested has proved unsuccessful.
Furthermore, decreasing the casing dn/t ratio, even as low as dn/t ¼ 6.0 in some cases, has not mitigated the problem. Furui et al. (2010)
performed a comprehensive investigation into the casing stability for a highly compacting chalk reservoir in the North Sea, concluding
that wells completed with a cluster perforating technique are prone to axial-compression collapsed-casing deformations. It was hypothe-
sized that the casing deformations are related to acid-injection volume and acid distribution over the productive interval and that a
“fully completed” well design should greatly reduce the risk of axial-compression collapse induced by cavity deformation.
Traditional well designs are cluster perforated, as shown below in Fig. 26a. The productive zone is perforated in 10-ft-long, 1-shot/ft
perforation clusters spaced at intervals 50–200 ft apart. Stimulation treatments are performed by bullheading 28% HCl acid and using
perforation ball sealers, coiled tubing, or limited entry to divert acid. Fig. 26b represents a fully completed well design where the
productive zone is perforated with 1 shot every 10–35 ft. Zone lengths are determined on the basis of the optimal-acid-volume and

602 April 2020 SPE Journal

ID: jaganm Time: 14:10 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119


J191625 DOI: 10.2118/191625-PA Date: 4-April-20 Stage: Page: 603 Total Pages: 22

pump-rate design guidelines, as discussed above. A limited-entry design with 800–1,200 psi of perforation friction is used to divert acid
along the treatment length. If the reservoir interval is too long to meet these design guidelines, then the acid-stimulation treatments are
pumped in stages with each previous stage being isolated by a retrievable bridge plug.

200 ft between 10 ft perf clusters


perf clusters with 1 SPF
(a) Cluster-perforated well with ball-sealer diversion bullhead acid treatment

(b) Fully completed well with limited-entry-diversion bullhead acid treatment

Fig. 26—Cluster-perforated well vs. fully completed well.

Data from North Sea fields have been reviewed to determine the effect of acid volume and diversion technique, on the wellbore
integrity between the two stimulation methods. Cluster-perforated wells have either 65/ 8-in., 65.8-lbm/ft, Q-125 (dn/t ¼ 5.9) or 51/ 2-in.,
32.6-lbm/ft, Q-125 (dn/t ¼ 8.8) casing to prevent pipe deformation. Fully completed wells have the same low D/t ratio casings or
75/ 8-in., 51.2-lbm/ft, Q-125 (dn/t ¼ 11.1) or 5-in., 23.2-lbm/ft, Q-125 (dn/t ¼ 10.5). Well records from intervention or wireline operations
were used to determine the minimum casing inside diameter (ID) on the basis of tool access. If casing minimum ID was less than
4.1 in., compared to the original casing minimum drift ID ¼ 4.25 in., it was classified as a deformation. In some cases, caliper data were
also available. The failure criterion was set to the minimum ID for a CT-straddle-tool outside diameter. On the basis of PLT data,
restimulation is required for most cluster-perforated wells and coiled tubing with straddle packers is recommended to achieve effective
diversion for the restimulation.
As shown in Fig. 27, more than 80% of cluster-perforated wells experienced pipe deformations in the reservoir liner when investi-
gated within 5 years of well life. Blue bars represent the number of wells investigated that have reached that age, and the red bars repre-
sent the number of failed wells detected at that age. Despite using thick liners with low dn/t ratios, pipe deformation has not been
prevented successfully. These cluster-perforated wells are typically acidized, reacidized, and scale squeezed using bullhead techniques.

Well-Failure Statistics in Cluster-Perforated Wells


50 100

45 90
42
40
40 38 80
35 36 36
35 33 33 33 70
Number of Wells

32
29 30
30 60
26 26 26
Failure (%)

25 50 Investigated wells
20 19 19 40 Failed wells
15 Failure % investigated wells
15 30

10 20
6
5 10
0
0 0
r rs rs rs rs rs rs rs rs rs
yea ea ea ea ea ea ea ea ea ea
<1 < 2y < 3y < 4y < 5y < 6y < 7y < 8y < 9y 1 0y
X X <
r< r< <X <X <X <X <X <X <X <X
ea ea ea
rs
ea
rs
ea
rs
ea
rs
ea
rs
ea
rs
ea
rs rs
0y 1y 2y 3y 4y 5y 6y 7y 8y yea
9
Age of Well

Fig. 27—Well-failure statistics in cluster-perforated wells.

As shown in Fig. 28 below, fully completed wells have not experienced any pipe deformations, even though in some cases a higher-
dn/t-ratio casing has been run. At the time of this analysis, 45 wells had been completed and stimulated with this fully completed
matrix-acidizing technique, 9 wells have had multifinger calipers run, 20 wells have had PLTs run, and the remaining 5 have had other
various wireline interventions. Because this was a new stimulation technique in the field, wells were investigated routinely early in their
life for signs of liner collapse. Because no indications of failure were seen, these interventions were performed less frequently. There
are fewer old wells in this data set because this technique has been more recently implemented.

April 2020 SPE Journal 603

ID: jaganm Time: 14:10 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119


J191625 DOI: 10.2118/191625-PA Date: 4-April-20 Stage: Page: 604 Total Pages: 22

Well-Failure Statistics in Fully Completed Wells


50 100

45 90

40 80

35 70
Number of Wells

30 60

Failure (%)
Investigated wells
25 50 Failed wells
20 40 Failure % investigated wells

15 15
30
10 9 20
6
5 10
3
0 0 0 0 1 0
0 0
r rs rs rs rs
yea ea ea ea ea
<1 2y 3y 4y 5y
r<
X X< <X
<
<X
<
<X
<
ea r< rs rs rs
0y ea ea ea ea
1y 2y 3y 4y
Age of Well

Fig. 28—Well-failure statistics in fully completed wells.

A comparison between the acid-injection volume pumped per injection station (sleeve, cluster, or perforation) and per foot of pay
interval was made between the two techniques, as shown in Fig. 29. Cluster-perforated wells are shown by squares and fully completed
wells are shown by circles. Well failures are red. Because fully-completed wells are new to the field, cluster-perforated wells were fur-
ther broken down by age to allow for a direct comparison of similarly aged wells.

Correlation of Acid-Injection Volume and Well-Failure Occurence


20
Failure in cluster-perforated wells
28% HCl Volume per ft of Pay (bbl/ft)

18
No failure (0–4 years) in cluster-perforated wells
16 No failure (≥5 years) in cluster-perforated wells
No failure (0–4 years) in fully completed wells
14

12

10

0
10 100 1,000
28% HCl Volume per Sleeve, Cluster or Perf (bbl/station)

Fig. 29—Acid-injection volume and well-failure occurrence.

As shown in Fig. 29, high volumes of acid pumped into the formation and/or uneven acid distribution (poor acid coverage) results in
a higher number of casing deformations. Well failures correlate to higher acid volume injected per station. The data proved a hypothesis
that large volumes of acid create a cavity behind pipe (e.g., 100 bbl of 28% HCl acid dissolves 139 ft3 of chalk matrix for / ¼ 35%)
and confirm the recommendation made by Furui et al. (2010) that fully completed well designs would reduce the occurrence of liner
collapses compared to cluster-perforated well designs.
These pipe deformations might not deteriorate production significantly during the first onset, but there is a significant impact and
consequence of these failures over the life of a well. For this field, several different workover and intervention jobs were cancelled as a
result of the liner deformations including
1. Water shutoffs with mechanical isolation of high-water-cut and high-pressure zones.
2. Production logging for reservoir management and restimulation-candidate selection.
3. Coiled-tubing restimulations with coiled-tubing straddle packers to divert acid into underperforming intervals identified from
PLT runs, attributed to poor initial fluid diversion during acid treatment. Historical P50 uplift from coiled-tubing restimulation is
18–66% of prerestimulation well test.
4. Milling scale in reservoir liner to gain full access for PLT reperforating and restimulation.

Conclusions
The following conclusions can be drawn from this work:
• Better understanding of wormhole propagation during matrix acid treatments allows us to focus on pump rate and acid-flow distribu-
tion along the treatment interval to improve stimulation design.

604 April 2020 SPE Journal

ID: jaganm Time: 14:10 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119


J191625 DOI: 10.2118/191625-PA Date: 4-April-20 Stage: Page: 605 Total Pages: 22

• A skin value of 4 is readily achievable in carbonate reservoirs through the use of high-rate matrix-acidizing designs. This was
reconfirmed by our field data and the data that were published in the past. A short interval is easier to be treated effectively by acid
than a long interval or an interval with a thief zone. In such intervals, limited-entry design is one of the solutions to distribute acid
more evenly across the interval, as used in the industry for many years.
• The negative acid skin resulting from wormhole propagation is relatively stable under production stresses.
• The wormhole-penetration model provided by Furui et al. (2012a, 2012b) successfully predicts matrix-acidizing processes in both
laboratory- and field-scale work.
• Small- and large-block laboratory tests have reconfirmed wormholing efficiency, which was discussed as a scale effect in
previous studies.
• Water PVbt and acid PVbt were very close. This result indicates that there is a preferential, higher-permeability flow path within the
overall PV for this core sample. This also reconfirms that a series of destructive acid-coreflooding tests can be replaced by non-
destructive laboratory tests such as the tracer tests suggested by Zakaria et al. (2015). Further study on this subject is encouraged so
that field cores can be used for other purposes in the future.
• Real-time CT-scan tests were performed. A wormhole propagation was captured partially by taking a series of 2D CT images. A fur-
ther modification is required to obtain a full picture of how a wormhole propagates in a linear core system.

Nomenclature
dn ¼ nominal pipe diameter, in.
k ¼ reservoir permeability, md
kpost-acid ¼ permeability in the acid-stimulation zone, md
PVbt,opt ¼ optimal pore volumes to breakthrough, dimensionless
re ¼ drainage reservoir radius, ft
rw ¼ wellbore radius, ft
rwa ¼ effective wellbore radius, ft
rwh ¼ effective wormhole-propagation distance, ft
S ¼ skin factor, dimensionless
Spost-acid ¼ post-acid-stimulation skin factor, dimensionless
Spre-acid ¼ pre-acid-stimulation skin factor, dimensionless
t ¼ pipe thickness, in.
Dt ¼ elapsed time, hours
vi ¼ interstitial velocity, cm/min
r ¼ external stress, psi
/ ¼ porosity, dimensionless

Acknowledgment
The authors would like to thank ConocoPhillips Company for permission to publish this paper.

References
Abou-Sayed, I. S., Shuchart, C. E., Choi, N. H. et al. 2007. Well Stimulation Technology for Thick, Middle East Carbonate Reservoirs. Paper presented
at the International Petroleum Technology Conference, Dubai, UAE, 4–6 December. IPTC-11660-MS. https://doi.org/10.2523/IPTC-11660-MS.
Aidagulov, G., Qiu, X., Brady, D. et al. 2018. New Insights into Carbonate Matrix Stimulation from High-Resolution 3D Images of Wormholes Obtained
in Radial Acidizing Experiments. Paper presented at the SPE Kingdom of Saudi Arabia Annual Technical Symposium and Exhibition, Dammam,
Saudi Arabia, 23–26 April. SPE-192366-MS. https://doi.org/10.2118/192366-MS.
Akanni, O. O. and Nasr-El-Din, H. A. 2015. The Accuracy of Carbonate Matrix-Acidizing Models in Predicting Optimum Injection and Wormhole Prop-
agation Rates. Paper presented at the SPE Middle East Oil & Gas Show and Conference, Manama, Bahrain, 8–11 March. SPE-172575-MS. https://
doi.org/10.2118/172575-MS.
Albuquerque, M. A., Ledergerber, A. G., Smith, C. L. et al. 2006. Use of Novel Acid System Improves Zonal Coverage of Stimulation Treatments
in Tengiz Field. Paper presented at the SPE International Symposium and Exhibition on Formation Damage Control, Lafayette, Louisiana, USA,
15–17 February. SPE-98221-MS. https://doi.org/10.2118/98221-MS.
Al-Dahlan, M. N. and Nasr-El-Din, H. A. 2000. A New Technique to Evaluate Matrix Acid Treatments in Carbonate Reservoirs. Paper presented at the SPE Inter-
national Symposium on Formation Damage Control, Lafayette, Louisiana, USA, 23–24 February. SPE-58714-MS. https://doi.org/10.2118/58714-MS.
Al-Dhafeeri, A. M., Engler, T. W., and Nasr-El-Din, H. A. 2002. Application of Two Methods to Evaluate Matrix Acidizing Using Real-Time Skin
Effect in Saudi Arabia. Paper presented at the SPE International Symposium and Exhibition on Formation Damage Control, Lafayette, Louisiana,
USA, 20–21 February. SPE-73703-MS. https://doi.org/10.2118/73703-MS.
Anvik, H. K. and Gibson, W. R. 1987. Drilling and Workover Experiences in the Greater Ekofisk Area. Paper presented at the SPE/IADC Drilling Con-
ference, New Orleans, Louisiana, USA, 15–18 March. SPE-16150-MS. https://doi.org/10.2118/16150-MS.
Arangath, R., Hopkins, K. W., Lungershausen, D. et al. 2008. Successful Stimulation of Thick, Naturally-Fractured Carbonates Pay Zones in Kazakhstan.
Paper presented at the SPE International Symposium and Exhibition on Formation Damage Control, Lafayette, Louisiana, USA, 13–15 February.
SPE-112419-MS. https://doi.org/10.2118/112419-MS.
Aslam, J. and Al Salat, T. 1998. Stimulation of Horizontal Wells in Carbonate Reservoirs. Paper presented at the 8th Abu Dhabi International Petroleum
Exhibition and Conference, Abu Dhabi, UAE, 11–14 October. SPE-49493-MS. https://doi.org/10.2118/49493-MS.
Barkved, O., Heavey, P., Kjelstadli, R., Kleppan, T., Kristiansen, T. G. 2003. Valhall Field—Still on Plateau after 20 Years of Production. Paper pre-
sented at the Offshore Europe 2003, Aberdeen, United Kingdom, 2–5 September. SPE-83957-MS. https://doi.org/10.2118/83957-MS.
Bartko, K. M., Acock, A. M., Robert, J. A. et al. 1997. A Field Validated Matrix Acidizing Simulator for Production Enhancement in Sandstone and Car-
bonates. Paper presented at the SPE European Formation Damage Conference, The Hague, The Netherlands, 2–3 June. SPE-38170-MS. https://
doi.org/10.2118/38170-MS.
Bazin, B. 2001. From Matrix Acidizing to Acid Fracturing: A Laboratory Evaluation of Acid/Rock Interactions. SPE Prod & Fac 16 (1): 22–29. SPE-
66566-PA. https://doi.org/10.2118/66566-PA.

April 2020 SPE Journal 605

ID: jaganm Time: 14:10 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119


J191625 DOI: 10.2118/191625-PA Date: 4-April-20 Stage: Page: 606 Total Pages: 22

Bazin, B., Bieber, M. T., Roque, C. et al. 1996. Improvement in Characterization of Acid Wormholing by In-Situ X-Ray CT Visualizations. Paper pre-
sented at the SPE Formation Damage Control Symposium, Lafayette, Louisiana, USA, 14–15 February. SPE-31073-MS. https://doi.org/10.2118/
31073-MS.
Bazin, B., Charbonnel, P., and Onaisi, A. 1999. Strategy Optimization for Matrix Treatments of Horizontal Drains in Carbonate Reservoirs, Use of Self-
Gelling Acid Diverter. Paper presented at the European Formation Damage Conference, The Hague, The Netherlands, 31 May–1 June. SPE-54720-MS.
https://doi.org/10.2118/54720-MS.
Bitanov, A. A., Nadezhdin, S. V., and Adilgereev, R. S. 2005. Productivity Impact from Matrix Stimulating Thick Intervals Using a Nondamaging Self-
Diverting Acidizing System: A Northwestern Kazakhstan Case Study. Paper presented at the 14th SPE Middle East Oil & Gas Show and Conference,
Bahrain, Saudi Arabia, 12–15 March. SPE-92831-MS. https://doi.org/10.2118/92831-MS.
Buijse, M. A. 2000. Understanding Wormholing Mechanism Can Improve Acid Treatments in Carbonate Formations. SPE Prod & Fac 15 (3): 168–175.
SPE-65068-PA. https://doi.org.10.2118/65068-PA.
Buijse, M. A. and Glasbergen, G. 2005. A Semianalytical Model to Calculate Wormhole Growth in Carbonate Acidizing. Paper presented at the SPE
Annual Technical Conference and Exhibition, Dallas, Texas, USA, 9–12 October. SPE-96892-MS. https://doi.org/10.2118/96982-MS.
Chambers, K. T., Hallager, W. S., Kabir, C. S. et al. 1997. Characterization of a Carbonate Reservoir Using Pressure Transient Tests and Production
Logs: Tengiz Field, Kazakhstan. Paper presented at the SPE Annual Technical Conference and Exhibition, San Antonio, Texas, USA, 5–8 October.
SPE-38657-MS. https://doi.org/10.2118/38657-MS.
Cheng, H., Zhu, D., and Hill, A. D. 2017. The Effect of CO2 on Wormhole Propagation in Carbonate Acidizing. SPE Prod & Oper 32 (3): 325–332.
SPE-178962-PA. https://doi.org/10.2118/178962-PA.
Clancey, B., Aly, M., Bugti, M. et al. 2014. Barzan Completions Success through Innovative Stimulation and Testing Technologies. Paper presented at
the International Petroleum Technology Conference, Doha, Qatar, 19–22 January. IPTC-17307-MS. https://doi.org/10.2523/IPTC-17307-MS.
Cui, B., Feng, P., Rong, X. et al. 2014. Acidizing Technology of Carbonate Reservoir Used for Enhanced Oil Recovery in the Oilfield of Iraq. Paper pre-
sented at the 24th International Ocean and Polar Engineering Conference, Busan, South Korea, 15–20 June. ISOPE-I-14-035.
Daccord, G., Touboul, E., Lenormand, R. 1989. Carbonate Acidizing: Toward a Quantitative Model of the Wormholing Phenomenon. SPE Prod Eng
4 (1): 63–68. SPE-16887-PA. https://doi.org/10.2118/16887-PA.
Dong, K., Jin, X., Zhu, D. et al. 2014. The Effect of Core Dimensions on the Optimum Acid Flux in Carbonate Acidizing. Paper presented at the SPE
International Symposium and Exhibition on Formation Damage Control, Lafayette, Louisiana, USA, 26–28 February. SPE-168146-MS. https://
doi.org/10.2118/168146-MS.
Economides, M. J., Hill, A. D., and Ehlig-Economides, C. 1994. Petroleum Production System, first edition. Upper Saddle River, New Jersey, USA:
Prentice Hall.
Foglio, G. and Wtorek, S. 2010. Kashagan Field: Defining Stimulation Methodology, Implementation, and Results Evaluation to Provide a Program
Redesigning Criterion. Paper presented at the SPE Annual Technical Conference and Exhibition, Florence, Italy, 19–22 September. SPE-134403-MS.
https://doi.org/10.2118/134403-MS.
Folomeev, A. E., Sharifullin, A. R., Murinov, K. Y. et al. 2014. Theory and Practice of Acidizing High Temperature Carbonate Reservoirs of R. Trebs
Oil Field, Timan-Pechora Basin. Paper presented at the SPE Russian Oil and Gas Exploration and Production Technical Conference and Exhibition,
Moscow, Russia, 14–16 October. SPE-171242-MS. https://doi.org/10.2118/171242-MS.
Fredd, C. N. and Fogler, H. S. 1996. Alternative Stimulation Fluids and Their Impact on Carbonate Acidizing. Paper presented at the SPE Formation
Damage Control Symposium, 14–15 February, Lafayette, Louisiana, USA. SPE-31074-MS. https://doi.org/10.2118/31074-MS.
Fredd, C. N. and Miller, M. J. 2000. Validation of Carbonate Matrix Stimulation Models. Paper presented at the SPE International Symposium on Forma-
tion Damage Conference, Lafayette, Louisiana, USA, 23–24 February. SPE-58713-MS. https://doi.org/10.2118/58713-MS.
Frick, T. P., Mostofizadeh, B., and Economides, M. J. 1994. Analysis of Radial Core Experiments for Hydrochloric Acid Interaction with Limestones.
Paper presented at the SPE Formation Damage Control Symposium, Lafayette, Louisiana, USA, 7–10 February. SPE-27402-MS. https://doi.org/
10.2118/27402-MS.
Furui, K, Burton, R. C., Burkhead, D. W. et al. 2012a. A Comprehensive Model of High-Rate Matrix-Acid Stimulation for Long Horizontal Wells in
Carbonate Reservoirs: Part I—Scaling Up Core-Level Acid Wormholing to Field Treatments. SPE J. 17 (1): 271–279. SPE-134265-PA. https://
doi.org/10.2118/134265-PA.
Furui, K. Burton, R. C., Burkhead, D. W. et al. 2012b. A Comprehensive Model of High-Rate Matrix-Acid Stimulation for Long Horizontal Wells in
Carbonate Reservoirs: Part II—Wellbore/Reservoir Coupled-Flow Modeling and Field Application. SPE J. 17 (1): 280–291. SPE-155497-PA.
https://doi.org/10.2118/155497-PA.
Furui, K., Fuh, G.-F., Abdelmalek, N. A. et al. 2010. Comprehensive Modeling Analysis of Borehole Stability and Production Liner Deformation for
Inclined/Horizontal Wells Completed in a Highly Compacting Chalk Formation. SPE Drill & Compl 25 (4): 530–543. SPE-123651-PA. https://
doi.org/10.2118/123651-PA.
Gdanski, R. 1999. A Fundamentally New Model of Acid Wormholing in Carbonates. Paper presented at the European Formation Damage Conference,
The Hague, The Netherlands, 31 May–1 June. SPE-54719-MS. https://doi.org/10.2118/54719-MS.
Ghommem, M., Qiu, X., Brady, D. et al. 2016. Monitoring of Matrix Acidizing by Using Resistivity Measurements. Presented at the SPE Annual Techni-
cal Conference and Exhibition, Dubai, UAE, 26–28 September. SPE-181414-MS. https://doi.org/10.2118/181414-MS.
Gil, I., Ramos, G. G., Montgomery, C. T. et al. 2005. Failure Mechanisms in Deep-Water Chalks: Rock Stability as Function of Effective Stress and
Saturation. 2005. Paper presented at the 40th U.S. Symposium on Rock Mechanics, Anchorage, Alaska, USA, 25–29 June. ARMA/USRMS 05-718.
Gong, M. and El-Rabaa, A. M. 1999. Quantitative Model of Wormholing Process in Carbonate Acidizing. Paper presented at the SPE Mid-Continent
Operations Symposium, Oklahoma City, Oklahoma, USA, 28–31 March. SPE-52165-MS. https://doi.org/10.2118/52165-MS.
Haldar, S., Al-Jandal, A. A., and Al-Driweesh, S. M. 2008. Evaluation of Rotary Jetting Tool Application for Matrix Acid Stimulation of Carbonate Res-
ervoir in Southern Area Field of Saudi Arabia. Paper presented at the International Technology Conference, Kuala Lumpur, Malaysia, 3–5 December.
IPTC-12023-MS. https://doi.org/10.2523/IPTC-12023-MS.
Hawkins, M. F. 1956. A Note on the Skin Effect. J Pet Technol 8 (12): 65–66. SPE-732-G. https://doi.org/10.2118/732-G.
Hoefner, M. L. and Fogler, H. S. 1988. Pore Evolution and Channel Formation During Flow and Reaction in Porous Media. AICheE J 34 (1): 45–54.
https://doi.org/10.1002/aic.690340107.
Hung, K. M., Hill, A. D., and Sepehrnoori, K. 1989. A Mechanistic Model of Wormhole Growth in Carbonate Matrix Acidizing and Acid Fracturing.
J Pet Technol 41 (1): 59–66. SPE-16886-PA. https://doi.org/10.2118/16886-PA.
Issa, A. A., Uematsu, H., and Bellah, S. 2015. Understanding and Overcoming Challenges of Carbonate Reservoir Matrix Stimulation—The Case of an
Offshore Abu Dhabi Reservoir. Paper presented at the Abu Dhabi International Petroleum Exhibition and Conference, Abu Dhabi, UAE,
9–12 November. SPE-177756-MS. https://doi.org/10.2118/177756-MS.

606 April 2020 SPE Journal

ID: jaganm Time: 14:10 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119


J191625 DOI: 10.2118/191625-PA Date: 4-April-20 Stage: Page: 607 Total Pages: 22

Izgec, O., Keys, R. S., Zhu, D. et al. 2008. An Integrated Theoretical and Experimental Study on the Effects of Multiscale Heterogeneities in Matrix
Acidizing of Carbonates. Paper presented at the Annual Technical Conference and Exhibition, Denver, Colorado, USA, 21–24 September. SPE-
115143-MS. https://doi.org/10.2118/115143-MS.
Jardim Neto, A. T., Silva, C. A. M., Torres, R. S. et al. 2013. Self-Diverting Acid for Effective Carbonate Stimulation Offshore Brazil: A Successful His-
tory. Paper presented at the SPE European Formation Damage Conference and Exhibition, Noordwijk, The Netherlands, 5–7 June. SPE-165089-MS.
https://doi.org/10.2118/165089-MS.
Karale, C., Beuterbaugh, A., Pinto, M. et al. 2016. HP/HT Carbonate Acidizing—Recent Discoveries and Contradictions in Wormhole Phenomenon. Paper
presented at the Offshore Technology Conference Asia, Kuala Lumpur, Malaysia, 22–25 March. OTC-26714-MS. https://doi.org/10.4043/26714-MS.
Kayumov, R., Konchenko, A., Bairamov, A. et al. 2014. Experience of Carbonate Acidizing in the Challenging Environment of the Volga-Urals Region
of Russia. Paper presented at the SPE International Symposium and Exhibition on Formation Damage Control, Lafayette, Louisiana, USA,
26–28 February. SPE-168167-MS. https://doi.org/10.2118/168167-MS.
Kazemi, H., Merrill, L. S., and Jargon, J. R. 1972. Problems in Interpretation of Pressure Fall-Off Tests in Reservoir with and without Fluid Banks.
J Pet Technol 24 (9): 1147–1156. SPE-3696-PA. https://10.2118/3696-PA.
Kent, A. W., Burkhead, D. W., Burton, R. C. et al. 2014. Intelligent Completion Inside Uncemented Liner for Selective High-Rate Carbonate Matrix
Acidizing. SPE Drill & Compl 29 (2): 165–181. SPE-166209-PA. https://doi.org/10.2118/166209-PA.
Lagrone, K. W. and Rasmussen, J. W. 1963. A New Development in Completion Methods—The Limited Entry Technique. J Pet Technol 15 (7):
695–702. SPE-530-PA. https://doi.org/10.2118/530-PA.
MaGee, J., Buijse, M. A., and Pongratz, R. 1997. Method for Effective Fluid Diversion when Performing a Matrix Acid Stimulation in Carbonate Forma-
tions. Paper presented at the Middle East Oil Show, Bahrain, Saudi Arabia, 15–18 March. SPE-37736-MS. https://doi.org/10.2118/37736-MS.
McDuff, D., Shuchart, C. E., Jackson, S. et al. 2010. Understanding Wormholes in Carbonates: Unprecedented Experimental Scale and 3-D Visualiza-
tion. Paper presented at the SPE Annual Technical Conference and Exhibition, Florence, Italy, 19–22 September. SPE-134379-MS. https://doi.org/
10.2118/134379-MS.
McLeod, H. O. and Coulter, A. W. 1969. The Stimulation Treatment Pressure Record—An Overlooked Formation Evaluation Tool. J Pet Techol 21 (8):
951–960. SPE-2287-PA. https://10.2118/2287-PA.
Montgomery, C. T., Ramos, G. G., Gil, I. et al. 2005. Failure Mechanisms in Deepwater Chalks: Rock Stability as Function of Pore Pressure and Satur-
ation. Paper presented at the International Petroleum Technology Conference, Doha, Qatar, 21–23 November. IPTC-10321-MS. https://doi.org/
10.2523/IPTC-10321-MS.
Morita, N., Doi, T., and Kinoshita, T. 2005. Stability of an Openhole Completed in a Limestone Reservoir with and without Acid Treatments. SPE J.
10 (2): 105–114. SPE-77776-PA. https://doi.org/10.2118/77776-PA.
Mostofizadeh, B. and Economides, M. J. 1994. Optimum Injection Rate from Radial Acidizing Experiments. Paper presented at the SPE Annual Techni-
cal Conference and Exhibition, New Orleans, Louisiana, USA, 25–28 September. SPE-28547-MS. https:// doi.org/10.2118/28547-MS.
Nozaki, M., Burton, R. C., Furui, K. et al. 2014. Review and Analysis of Zonal Isolation Effectiveness in Carbonate Reservoirs Using Multi-Stage Stimu-
lation Systems. Paper presented at the SPE Annual Technical Conference and Exhibition, Houston, Texas, USA, 28–30 September. SPE-174755-MS.
https://doi.org/10.2118/174755-MS.
Paccaloni, G., Tabmbini, M., and Galoppini, M. 1988. Key Factors for Enhanced Results of Matrix Stimulation Treatments. Paper presented at the SPE
Formation Damage Control Symposium, Bakersfield, California, USA, 8–9 February. SPE-17154-MS. https://doi.org/10.2118/17154-MS.
Paccaloni, G. and Tambini, M. 1993. Advances in Matrix Stimulation Technology. J Pet Technol 45 (3): 256–263. SPE-20623-PA. https://doi.org/
10.2118/20623-PA.
Pichler, T., Frick, T. P., and Economides, M. J. 1992. Stochastic Modeling of Wormhole Growth in Carbonate Acidizing with Biased Randomness.
Paper presented at the European Petroleum Conference, Cannes, France, 16–18 November. SPE-25004-MS. https://doi.org/10.2118/25004-MS.
Prouvost, L. P. and Economides, M. J. 1987. Real-Time Evaluation of Matrix Acidizing Treatments. J Pet Sci Eng 1 (2): 145–154. https://doi.org/
10.1016/0920-4105(87)90005-2.
Prouvost, L. P. and Economides, M. J. 1989. Applications of Real-Time Matrix-Acidizing Evaluation Method. SPE Prod Eng 4 (4): 401–407. SPE-
17156-PA. https://doi.org/10.2118/17156-PA.
Qiu, X., Zhao, W., Chang, F. et al. 2013. Quantitative Modeling of Acid Wormholing in Carbonates—What Are the Gaps to Bridge. Paper presented at
the SPE Middle East Oil and Gas Show and Conference, Manama, Bahrain, 10–13 March. SPE-164245-MS. https://doi.org/10.2118/164245-MS.
Qiu, X., Aidagulov G., Ghommem, M. et al. 2018. Experimental Investigation of Radial and Linear Acid Injection into Carbonates for Well Stimulation
Operations. Paper presented at the SPE Kingdom of Saudi Arabia Annual Technical Symposium and Exhibition, Dammam, Saudi Arabia,
23–26 April. SPE-192261-MS. https://doi.org/10.2118/192261-MS.
Rachmawati, F. D., Andika, R., Simanjuntak, T. M. et al. 2008. Matrix Stimulation of Thick Carbonate Formation: High Rate Acid and CT Acid
Treatments—A Case History. Paper presented at the Abu Dhabi International Petroleum Exhibition and Conference, Abu Dhabi, UAE, 3–6 November.
SPE-117145-MS. https://doi.org/10.2118/117145-MS.
Sannier, Y., de Lapasse, H., Kirkhus, R. et al. 1999. Boosting Horizontal Well Performance in Carbonates by Selective Stimulation: Field Cases from
Qatar. Paper presented at the SPE European Formation Damage Conference, The Hague, The Netherlands, 31 May–1 June. SPE-54739-MS. https://
doi.org/10.2118/54739-MS.
Schechter, R. S. 1992. Oil Well Stimulation. Englewood Cliffs, New Jersey, USA: Prentice Hall.
Schwalbert, M. P., Zhu, D., and Hill, A. D. 2017. Extension of an Empirical Wormhole Model for Carbonate Matrix Acidizing through Two-Scale Con-
tinuum 3D Simulations. Paper presented at the SPE Europec featured at 79th EAGE Conference and Exhibition, Paris, France, 12–15 June. SPE-
185788-MS. https://doi.org/10.2118/185788-MS.
Sharag, M., El-Enien, S. A., Raya, O. A. et al. 2000. A Novel Approach to Acidizing of Carbonates: Case History from the Gulf of Suez in Egypt. Paper pre-
sented at the SPE Annual Technical Conference and Exhibition, Dallas, Texas, USA, 1–4 October. SPE-63182-MS. https://doi.org/10.2118/63182-MS.
Siddiqui, S., Nasr-El-Din, H. A., and Khamees, A. A. 2006. Wormhole Initiation and Propagation of Emulsified Acid in Carbonate Cores Using Compu-
terized Tomography. J Pet Sci Eng 54: 93–111. https://doi.org/10.1016/j.petrol.2006.08.005.
Singh, J. R. 1985. An Evaluation of Acid Frac/Matrix Stimulation of a Tight Limestone Formation in Exploratory Wells in Kuwait. Paper presented at
the Middle East Oil Technical Conference and Exhibition, Bahrain, Saudi Arabia, 11–14 March. SPE-13751-MS. https://doi.org/10.2118/13751-MS.
Tardy, P. M. J., Lecerf, B., and Christanti, Y. 2007. An Experimentally Validated Wormhole Model for Self-Diverting and Conventional Acids in Car-
bonate Rocks under Radial Flow Conditions. Paper presented at the European Formation Damage Conferences, Scheveningen, The Netherlands,
30 May–1 June. SPE-107854-MS. https://doi.org/10.2118/107854-MS.
Thabet, S., Brady, M., Parsons, C. et al. 2009. Changing the Game in the Stimulation of Thick Carbonate Gas Reservoirs. Paper presented at the Interna-
tional Petroleum Technology Conference, Doha, Qatar, 7–9 December. IPTC-13097-MS. https://doi.org/10.2523/IPTC-13097-MS.

April 2020 SPE Journal 607

ID: jaganm Time: 14:10 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119


J191625 DOI: 10.2118/191625-PA Date: 4-April-20 Stage: Page: 608 Total Pages: 22

Thomas, R. L. and Nasr-El-Din, H. A. 2003. Field Validation of a Carbonate Matrix Acidizing Model: A Case Study of Seawater Injection Wells in
Saudi Arabia. Paper presented at the SPE European Formation Damage Conference, The Hague, Netherlands, 13–14 May. SPE-82271-MS. https://
doi.org/10.2118/82271-MS.
Ussenbayeva, K. Utebaeva, D., Molesworth, G. et al. 2012. Successful Application of a Fit-for-Purpose Acid Program in the Tengiz Field. Paper pre-
sented at the Abu Dhabi International Petroleum Exhibition and Conference, Abu Dhabi, UAE, 11–14 November. SPE-160804-MS. https://doi.org/
10.2118/160804-MS.
Van Domelen, M. S., Hammoud, M. A., and Glasbergen, G. 2012. Optimization of Limited Entry Matrix Acid Stimulations with Laboratory Testing
and Treatment Pressure Matching. Paper presented at the SPE International Production and Operations Conference and Exhibition, Doha, Qatar,
14–16 May. SPE-157429-MS. https://doi.org/10.2118/157429-MS.
Walle, L. E. and Papamichos, E. 2015. Acidizing of Hollow Cylinder Chalk Specimens and Its Impact on Rock Strength and Wormhole Network Structure.
Paper presented at the 49th U.S. Rock Mechanics/Geomechanics Symposium, San Francisco, California, USA, 28 June–1 July. ARMA-2015-566.
Wang, H. and Schechter, R. S. 1993. The Optimum Injection Rate for Matrix Acidizing of Carbonate Formations. Paper presented at the SPE Annual
Technical Conference and Exhibition, Houston, Texas, USA, 3–6 October. SPE-26578-MS. https://doi.org/10.2118/26578-MS.
Williams, B. B., Gidley, J. L., and Schechter, R. S. 1979. Acidizing Fundamentals, Vol. 6. New York, New York, USA: Henry L. Doherty Series, Society
of Petroleum Engineers.
Yudovich, A., Chin, L. Y., and Morgan, D. R. 1989. Casing Deformation in Ekofisk. J Pet Technol 41 (7): 729–734. SPE-17856-PA. https://doi.org/
10.2118/17856-PA.
Zakaria, A. S., Nasr-El-Din, H. A., and Ziauddin, M. 2015. Predicting the Performance of the Acid-Stimulation Treatments in Carbonate Reservoirs with
Nondestructive Tracer Tests. SPE J. 20 (6): 1238–1253. SPE-174084-PA. https://doi.org/10.2118/174084-PA.
Zhu, D. and Hill, A. D. 1996. Real-Time Monitoring of Matrix Acidizing Including the Effects of Diverting Agents. SPE Prod & Fac 11 (2): 95–101.
SPE-28548-PA. https://doi.org/10.2118/28548-PA.

Robert C. Burton is a senior completions engineering fellow in the ConocoPhillips Global Drilling and Completions Group, special-
izing in well completions. He is a registered engineer in the states of California and Texas and has authored or coauthored a
number of SPE papers dealing with drilling and completions. Burton holds a bachelor’s degree in chemical engineering from the
Georgia Institute of Technology and a master’s degree in petroleum engineering from the University of Southern California. He is
an SPE Distinguished Member.
Manabu Nozaki is a senior completions engineer at ConocoPhillips Norway. His research interests include formation-failure analy-
sis, tubular-stability analysis, inflow-performance evaluation, and well-stimulation design and evaluation. Nozaki holds a BE
degree in resources and environmental engineering from Waseda University, Japan, and MS and PhD degrees in petroleum
engineering from Texas A&M University.
Nola R. Zwarich is a staff completions engineer at ConocoPhillips Alaska. She specializes in carbonate acid stimulation and also
has expertise in advanced tubular design, subsea completions, and intelligent well systems. Zwarich holds a BSc degree in
mechanical engineering, with a minor in petroleum engineering, from the University of Calgary.
Kenji Furui is an associate professor at Waseda University, Japan. His current research interests include rock mechanics, sand
control, borehole- and casing-stability analysis, well stimulation, and well-performance evaluation. Furui holds MS and PhD
degrees in petroleum engineering from the University of Texas at Austin. He is the recipient of the 2013 SPE Cedric K.
Ferguson Medal.

608 April 2020 SPE Journal

ID: jaganm Time: 14:10 I Path: S:/J###/Vol00000/190119/Comp/APPFile/SA-J###190119

You might also like