Download as pdf or txt
Download as pdf or txt
You are on page 1of 49

Distributed Fiber Optic Sensing System for Well-based Monitoring Water

Injection Tests—A Geomechanical Responses Perspective

Yankun Sun1, 2, Ziqiu Xue1, 2, Tsutomu Hashimoto1, 2, Xinglin Lei3, Yi Zhang1, 2

1
Geological Carbon Dioxide Storage Technology Research Association, 9-2 Kizugawadai, Kizugawa-shi, Kyoto 619-0292,

Japan

2
Research Institute of Innovative Technology for the Earth, 9-2 Kizugawadai, Kizugawa-shi, Kyoto 619-0292, Japan

3
Geological Survey of Japan, National Institute of Advanced Industrial Science and Technology (AIST), Central #7, Higashi

1-1-1, Tsukuba, Ibaraki 305-8567, Japan

Corresponding author: Yankun Sun (sun@rite.or.jp).

Key Points:

• Injection-induced geomechanical deformation (i.e., strain) induced by water injection is

continuously monitored via an advanced DFOS tool.

• Strain response mainly attributed to the injection scheme and lithological

heterogeneity agreed well with well logging and numerical results.

• DFOS-based downhole monitoring is a cost-effective tool to detect impacted zones and

quantify real-time wellbore deformation in depth.

Abstract

In the study, distributed fiber optic sensing (DFOS) based on hybrid Brillouin–Rayleigh

This article has been accepted for publication and undergone full peer review but has not
been through the copyediting, typesetting, pagination and proofreading process which may
lead to differences between this version and the Version of Record. Please cite this article as
doi: 10.1029/2019WR024794

©2019 American Geophysical Union. All rights reserved.


backscattering is examined for the first time in well-based monitoring distributed profiles of

geomechanical deformation induced by water injection. Field water injection tests are

conducted under different injection scenarios between an injection well (230 m, IW #2) and

5.5 m away from a fibered monitoring well (300 m, MW #1) by deploying cables behind the

casing at Mobara, Japan. Effects of injection rate, pressure, and lithological heterogeneity on

the geomechanical deformation are quantitatively monitored via a DFOS approach and

indicate that induced strains significantly depend on injection rate and pressure. The results

also indicate that DFOS results are reasonably consistent with simultaneous geophysical well

logging data and corresponding numerical simulation. The extent of impacted areas and

magnitude of near-wellbore strains are explored to evaluate formation heterogeneity and

fluid migration behaviors. The field testing of hybrid DFOS technology is expected to

definitely advance elaborate monitoring and wider applications, such as CO2

geosequestration sites.

Plain Language Summary

Geological CO2 storage (GCS) in deep saline aquifers is broadly recognized as possessing the

potential to play a key role in mitigating anthropogenic climate change. Valid monitoring

methods are important to characterize the spatial distribution of sequestered CO2 in

underground reservoirs. In the study, hybrid Brillouin–Rayleigh scattering based on

high-resolution distributed fiber optic sensing (DFOS) as an advanced monitoring tool to

measure the distribution of temperature or strain is deployed behind casing to a depth of

300 m in an actual field of MW #1 to detect the vertical profile of strain changes at various

©2019 American Geophysical Union. All rights reserved.


locations along the entire cable length. Water injection tests are implemented to inject

water from adjacent injection well IW #2 (approximately 5.5 m distant) toward the fibered

MW #1. Results indicate that the impacted zones induced by water injection are expanded

to the vertical formations near the fibered well, thereby assessing overlying and overlying

sealing. Additionally, the induced strain magnitude and state are largely associated with the

injection rate, pressure, and formation heterogeneity compare well with the results of well

logging and numerical modeling. The DFOS-based downhole monitoring technology can

capture continuous depth-based geomechanical deformation, which can serve as a valid

indicator to track the migration of injected fluids and act as an early-warning for potential

fluid leakage.

1. Introduction

Fluid injection and storage in deep reservoir formations, such as geological CO2 storage (GCS)

sites, involve complicated geomechanical processes. White et al. (2014) and Zoback et al.

(2012) showed that massive fluid injection can cause pore pressure increase, and in turn

potentially lead to geomechanical deformation, formation fracturing, fault reactivation,

microseismicity, and surface uplift among others as shown by Verdon et al. (2013) and

Rutqvist et al. (2014). Based on the mechanism, it is readily speculated that underground

fluid behaviors can be imaged via monitoring injection-induced geomechanical responses

(e.g., strain, temperature, and pressure).

Therefore, it is extremely important to obtain high-accuracy monitoring tools and

©2019 American Geophysical Union. All rights reserved.


instrumentation to collect real-time signals from fluid migration and spreading in multiple

subsurface environments (Shirzaei et al., 2016). Despite important advances in developing

and utilizing subsurface monitoring tools with the view to detect fluid flow, a few challenges

remain poorly addressed, such as the effect of high-pressure and high-temperature (HPHT)

conditions, uncertainty, limited visibility of reservoir formations (Baldwin, 2014; Kersey,

2000), lack of long-term in-situ tools (Shanafield et al., 2018), and low-resolution point data.

In this case, DFOS technology is increasingly in demand in the oil and gas industry and is

deemed as an ideal sensing solution to detect fluid migration and evolution in-situ to benefit

from its electromagnetic interference (EMI) immunity, long-lasting durability, installation

ease, and small physical footprint when compared with that of traditional electrical sensing

and foil strain gauge. Among various DFOS technologies, Brillouin optical time-domain

reflectometry (BOTDR), optical time-domain reflectometry (OTDR), and optical frequency

domain reflectometry (OFDR) are frequently used to simultaneously measure parameters

including strain, temperature, and pressure along the entire fiber length (Kogure & Okuda,

2018; Luca Schenato, 2017). Both are superior to other optical sensors, such as

quasi-distributed fiber Bragg grating (FBG) that is used for local measurement and

Raman-based distributed temperature sensing (DTS) that is used only for temperature

(Arnon et al., 2014; Sun et al., 2016; Sun, et al, 2017a, b; Tyler et al., 2010). Although the DTS

system was marketed during the 1980s (Dakin et al., 1985; Kurashima et al., 1990) and is

used for well monitoring (Bakku et al., 2014; Duguid et al., 2017), it is still insufficient to view

complex fluid behaviors by solely relying on temperature sensing.

©2019 American Geophysical Union. All rights reserved.


More recently, an advanced DFOS system integrating Brillouin and Rayleigh scattering was

developed by Neubrex Co. Ltd., Japan (Kishida et al., 2012, 2014), and applied to strain

measurements associated with geomechanical research fields. Xue et al. (2018) and Zhang et

al. (2018) attached a fiber to Berea and Tako sandstone cores and proved the validity of

hybrid DFOS to monitor induced strain at varying confining/pore pressures. For field-scale

measurements, Xue et al. (2014), Xue & Hashimoto (2017), and Sun et al. (2018) deployed

the DFOS behind a well casing to sense geomechanical deformation in water injection,

extraction, and jet tests. Furthermore, the in-situ behaviors of micro-seismicity and

deformation are monitored via fiber optics in laboratories and/or field tests (Cappa et al.,

2006, 2008, 2019; Freifeld et al., 2014; Guglielmi et al., 2015; Moore et al., 2010; Sanada et

al., 2012). Despite the aforementioned recent tests, there is a paucity of field tests of fluid

injection and migration in the deep subsurface that measure the distributed geomechanical

deformation (i.e., strain) responses using DFOS. Hence, we explored capturing interwell

signals from water injection and migration in the subsurface using DFOS and examined

potential reasons for different induced strain responses in detail.

The main purpose of the present study is to experimentally implement and validate the

DFOS tool on-site based on hybrid Brillouin–Rayleigh scattering to vertically sense

geomechanical deformation following water injection between two adjacent wells in an

analogy of GCS sites. Water injection tests in sections ① and ②aare performed to

investigate the effects of injection rate, pressure, and lithological heterogeneity on the

optical response. Testing results demonstrate that impacted zone strain profiles are closely

©2019 American Geophysical Union. All rights reserved.


related to the injection scenario and lithological heterogeneity as evidenced by resistivity

logging and numerical simulation. Thus, the study is a meaningful and innovative field trial

to view fluid evolution using DFOS, which can provide continuous strain profiles in depth to

assess reservoir–caprock integrity and storage security. Additionally, it can guide

geomechanical engineers with in-depth knowledge and specifications on decision-making

for future GCS-related field applications and DFOS system deployment.

2 Methodology

2.1 Distributed Fiber Optic Sensing Technology

Based on the changes in the fiber optical properties (such as wavelength, frequency, and

intensity), DFOS is an optical measurement device employed for measuring the optical

properties of a fiber as a function of location. From the optical properties of the fiber,

magnitudes and locations for a wide range of different measurands that affect the fiber can

be estimated. More specifically, the frequency or wavelength shifts of backscattering light

are directly related to the applied strain, temperature, and pressure changes along the fiber

length within a certain sampling interval.

The distributed sensors are mainly based on Rayleigh, Raman and Brillouin scattering, which

allow the detection of strain and/or temperature along the entire fiber and cover a distance

of tens of kilometers (Hartog, 2017). The distributed scattering diagram of interest is shown

in Figure S1. Raman scattering is widely used in DTS application because it simply senses

temperature change. The typical spatial resolution of DTS is approximately 1 m, and its

©2019 American Geophysical Union. All rights reserved.


performance rapidly degrades with decreases in the temperature. Conversely, Brillouin and

Rayleigh backscattering are relatively sensitive to pressure, temperature, and strain and are

sufficient to satisfy distributed measurement needs, although single-point FBG sensors

exhibit higher resolution and sensitivity to pressure response (Liu et al., 2007; L Schenato et

al., 2019).

2.2 Measuring Strain and Temperature via Hybrid Brillouin–Rayleigh System

The advanced DFOS system used in the study is a fully distributed hybrid Brillouin–Rayleigh

sensing system that does not require gratings to be written on the optical fiber. It consists of

a high spatial resolution synthetic BOTDR (S-BOTDR) and a high-resolution Tunable

Wavelength Coherent Optical Time-Domain Reflectometry (TW-COTDR) technique, thereby

allowing the simultaneous measurement of Brillouin and Rayleigh scattering. Specifically, the

TW-COTDR can achieve the same sensitivity as that of a conventional FBG system

(Delepine-Lesoille et al., 2013) and is particularly well suited to detect small temperature

and/or strain changes. When strain changes occur at a certain location, the scattered signal

is modulated at that location. Strain variation locations can be precisely determined by

measuring the light time of flight. Furthermore, the system obtains a higher spatial

resolution of 2 cm when compared with that of 1 m resolution for conventional optical

interrogators.

The relationship between the Brillouin frequency shift (∆𝜐𝐵 ), strain variation (𝛥𝜀), and

temperature variation (𝛥𝑇) is given as follows:

∆𝜐𝐵 =𝐶11 Δ𝜀+𝐶12 Δ𝑇 (1)

©2019 American Geophysical Union. All rights reserved.


where 𝐶11 denotes the strain coefficient and 𝐶12 denotes the temperature coefficient for

the Brillouin frequency shift.

The Rayleigh frequency difference is obtained by comparing the measured value to the

reference state via cross-correlation. The Rayleigh frequency shift (∆𝜐𝑅 ) caused by strain

variations (Δ𝜀) and/or temperature variations (Δ𝑇) is expressed as follows:

∆𝜐𝑅 =𝐶21 Δ𝜀+𝐶22 Δ𝑇 (2)

where 𝐶21 denotes the strain coefficient and 𝐶22 denotes the temperature coefficient for

the Rayleigh frequency shift.

For a standard single-mode fiber, the hybrid system provides frequency shifts for Brillouin

and Rayleigh scatterings. As both shifts are a function of strain and temperature, their

separation is required. This removes the effect of temperature on strain, and vice versa, and

allows one to obtain pure strain and temperature values.

By combining equations (1) and (2), the strain and temperature are separated as follows:

𝛥𝜈𝐵 𝐶 𝐶12 𝛥𝜀 𝐶 𝐷12 𝛥𝜀 𝑡𝑜𝑡𝑎𝑙


[ ] = [ 11 ] [ ] = [ 11 ][ ] (3)
𝛥𝜈𝑅 𝐶21 𝐶22 𝛥𝑇 𝐶21 𝐷22 𝛥𝑇
where

𝐶12 = 𝐷12 + 𝛼𝐶11 (4)

𝐶22 = 𝐷22 + 𝛼𝐶21 (5)

and

𝛥𝜀 𝑡𝑜𝑡𝑎𝑙 = 𝛥𝜀 + 𝛼𝛥𝑇 (6)

In equations (4)–(6), 𝛼 denotes the thermal expansion coefficient. Equation (6) represents

©2019 American Geophysical Union. All rights reserved.


the total strain (𝛥𝜀 𝑡𝑜𝑡𝑎𝑙 ), which is a sum of its mechanical and thermal components. In

addition, Yamauchi (2010) provided an effective method to determine the coefficients for

the selected coated fiber.

3 Field Testing

3.1 Site Description

The field site is in northwestern Mobara city, Chiba Pref., Japan (boxed in cyan in Figure 1 (a)).

The sedimentary succession of the site mainly consists of silt and sand-silt in the hilly part of

the Kazusa Group and the Kanto Loam Formation and the sand layer below the Shimousasa

Subtropical Basin (Fukuda et al., 2015). From the geologic profile of the fiber-installed well

(see Figure S2), it is observed that the sediments of the test site are mainly from the surface

to the 300-m depth, Kasamori Formation, Chonan Formation (~120 m thick) and Kakinokidai

Formation (154 m in thickness) and belong to the upper Kazusa group. High-pressure water

is injected into the Triassic Formation at approximately 194–230 m depth.

3.2 Configurations of Two Adjacent Wells

In the study, two adjacent wells with depths of 230 m and 300 m were vertically drilled to

implement the water injection tests, as shown in Figures 1(b), (c), and 3. The 230-m deep

well acting as an injection well (IW #2) was designed subsequent to the 300-m deep well

monitoring well (MW #1). Additionally, the two wells were 5.5 m apart such that the

migrating zone of injected water was situated in the same formation (i.e., the sand layer).

Water was injected into a target layer connecting both wells to verify the responses of

©2019 American Geophysical Union. All rights reserved.


optical fibers and deformation resulting from the water injection. Three optical fiber cables

of a different design (different outer jackets) were cemented within the annulus between

the casing and formation wall (Xue & Hashimoto, 2017). Additionally, the pressure and

temperature of the injected water were also detected via pressure sensors built into the

drilling pipe near the water injection ports between the two packers (Figure 1d).

The configuration properties of the water injection performed in IW #2 are listed in Table 1.

It is important to recognize that the geomechanical deformation (i.e., strain) induced by the

water injection is relatively small, and thus, it is difficult to record Brillouin scattering mainly

due to its low sensitivity. A single-layer armored (1A) cable, a three-layer armored (3A) cable

with a built-in fiber in metal tube (FIMT) in the second layer for temperature measurement

that is immune to strain and pressure, and an optical communication fiber (C) are integrated

into the fiber cables to compare the fiber sensitivity as required (see Figure S3). Their

coatings are identical and are composed of polyamide (-20~+85.5 ℃), and a stainless steel

wire is externally wound around these two types of armored cables to protect the central

optical fiber. It should be noted that all the data in the research are collected from 1A cable.

Furthermore, 3A and 1A cables are specially developed for downhole installation behind the

well casing and designed for logging tools in the tubing. Thus, C cable is treated as a

reference because it is widely used for telecommunication and fully examined in laboratory

tests to data. Specifically, C cable was not used to acquire data during the two water

injections. Alternatively, 1A cable exhibits the same structure as the core of 3A cable with

lower sensitivity and is used to investigate the effect of the armored coating.

©2019 American Geophysical Union. All rights reserved.


Two injection schemes were implemented in IW #2 as summarized in Table 1. The

geomechanical deformations and temperature changes resulting from water injection were

obtained via a single NBX-8000 interrogator produced by Neubrex Co., Ltd., Kobe, Japan. The

hybrid DFOS tool can sense small strain and temperature changes at high measurement

accuracy (10 με/ 0.5 ℃) and spatial resolution (5 cm) over a long-range distance (~25 km).

The accuracy of the DFOS technique was demonstrated in several previous laboratory and

field tests (Kogure et al., 2015; Kogure & Okuda, 2018; Xue et al., 2014; Xue & Hashimoto,

2017; Zhang et al., 2019). In the study, two water injection tests at sections ① (186.8–188.8

m) and ② (204.4–206.4 m) were presented and considered.

For a better understanding of the effect of the lithological heterogeneity of the reservoir

formation on the fiber monitoring, the raw cores drilled in MW #1 from the 235.5to 290.25

m depth were collected and imaged via a medical X-ray computed tomography (CT) scanner

(Aquilion ONE TSX 301A, Toshiba Medical Systems Corp.), and core photographs and CT

images corresponding to the perforated interval (275.5–280.5 m) were also obtained. The

whole 25 m-long core CT images also manifested the formation skeleton (rock matrix) that is

relatively uniform and this is beneficial to cable installation and CO2 storage security.

Furthermore, a further analysis of the formation components based on gamma-ray logs

show that the sand–shale ratio is an important indication of injection-induced strain

magnitude, which is approximated as a function of the reservoir rock permeability and

porosity. Therefore, it is extremely important to characterize the formation heterogeneity via

CT scanning of the drilled cores and/or geophysical well logging in MW #1.

©2019 American Geophysical Union. All rights reserved.


Ultimately, fiber cables are deployed throughout the entire depth behind the well casing to

realize real-time monitoring of continuous strain/deformation profiles of the fullbore section.

In this case, only one cable can potentially link fluid injection into the surface deformation,

thereby allowing for the effects of lithologic heterogeneities and anomalies, which are

always overlooked or simplified in numerical simulations (Bachmann et al., 2012; Bissell et

al., 2011) or other surface-based monitoring methods, such as interferometric synthetic

aperture radar (InSAR) (Ringrose et al., 2009).

3.3 Field Test Procedure

Prior to the water injection, the wellbore and formations were equilibrated for a long period

of time (25 days). Two injection tests were performed at sections ① and ② and are

discussed more thoroughly in the subsequent subsections.

It is important to mention that for every injection test, Rayleigh and Brillouin signals were

concurrently logged utilizing this hybrid sensing interrogator to easily separate the

respective strain and temperature effects of Rayleigh signals from that of the Brillouin data,

although only TW-COTDR results were discussed in detail in the study. Another primary

reason is that actual monitoring results further revealed that Rayleigh scattering exhibits

better accuracy and higher signal-to-noise ratio (SNR) than S-BOTDR, as previously proven in

a few investigations (Nishiguchi, 2017; Nishiguchi et al., 2014). The resulting comparisons

between Rayleigh and Brillouin signals in the two field water injection tests are listed in the

Supporting Information section (see Figures S4, S5 and S6). Furthermore, multiple

geophysical well logging including electrical micro imager (EMI), radial cement bond log

©2019 American Geophysical Union. All rights reserved.


(RCBL), ultrasonic casing logs (𝑣𝑃 + 𝑣𝑆 ), pulsed neutron-gamma (GR + NL) logging, and

temperature measurement along the entire length of MW #1 were in progress in parallel

with the DFOS measuring before and/or during the water injection tests.

4 Results

4.1 Water Injection Tests at Section ①

The injection rate and pressure histories were logged by the sensor installed at 187.8 m in

IW #2 (Figure 3) throughout the water injection test. First, a small injection rate of 15 L/min

was used to inspect whether the induced deformation occurred at 10:04. However, given

the driving of the preexisting air, the water test was stopped for 27 min and restarted by

increasing the injection rate to 120 L/min at 10:56 and instantly reducing it to 80 L/min until

14:05. Additionally, pre-injected water was stored in the water tank, and its temperature

was approximately 12 °C, which was near that of the formation temperature. Hence, the

temperature effect recorded by the FIMT fiber on the geomechanical deformation was very

slight (see Figure S7) such that its results were ignored in the study.

Based on the data collected from the pressure sensor, injection pressure increased to

approximately 1.7 MPa before the injection start and immediately slightly increased slightly

once the test started. It was approximately 2.2 MPa after the injection restarted and

ultimately slightly decreased to approximately 2.1 MPa until the injection ended at 14:05,

thereby corresponding to the state of the injection rate. The cumulative injection volume

was estimated as approximately 17,500 L over 266 min.

©2019 American Geophysical Union. All rights reserved.


Time-varying frequency shifts recorded by a one-layer armored cable and in response to

converted strain are shown in Figure 4. After 228.3 min of injection time, the frequency

shifts increased to –9 GHz converted into +60 με strain, thereby implying that the fiber

cable was subjected to an expansive state as observed in a few previous studies (Shirzaei et

al., 2016; Verdon et al., 2013; Xue & Hashimoto, 2017). Furthermore, the first strain trough

at 184.2 m was owing to the occurrence of a clamp for fixing fiber cables to maintain their

straightness. The second at 189.2 m was mainly due to the coupling protector used for

protecting fiber cables that smoothly crossed the casing joints.

In Figure 5, the geomechanical deformation (i.e., strain) profile due to water injection was

plotted based on the data from the fiber cables in MW #1. The Rayleigh frequency shift

sampled at every 10 min (exactly 10.3 min) with the 1A cable was converted into strain and

displayed as a 2D contour and compared with an EMI image and resistivity map. In addition,

as previously stated, there is no significant observed temperature change during water

injection and thus, temperature compensation was not desirable in terms of the measured

results.

The maximum vertical extension of the impacted zone was mainly centered at 185–191 m.

This is consistent with their strain responses at the fifth time stamp as shown by the red

dashed line in Figure 4. The strain gradually increased by up to 60 microstrain during the

water injection.

4.2 Water Injection Test at Section ②

©2019 American Geophysical Union. All rights reserved.


For section ②, the injection rate and pressure histories with a time lapse are measured

employing a pressure sensor installed at 205.4 m in IW #2, as shown in Figure 6. The test

was performed with injection rates corresponding to 150 L/min (14 min), 60 L/min (78 min),

and 90 L/min (52 min) in sequence. Additionally, at 13:14, the injection pressure increased

by 0.13 MPa (ΔP = P(t=13:14) −P0 = 2.02 − 1.89 = 0.13 MPa) and the injection rate subsequently

increased to 150 L/min. At 10:44 for section ①, the injection pressure increased by 0.55

MPa (ΔP = P(t=10:44) −P0 = 2.17 − 1.62 = 0.55 MPa) and 𝑉i increased to 120 L/min. Thus, it is

inferred that the permeability of the injection section ② exceeded that of section ①.

Meanwhile, the injection pressure was maintained at 1.9 MPa before the test and then

gradually increased with time and reached a maximum of 2.1 MPa; thus, the pressure

increment was approximately 0.2 MPa. The cumulative injection volume was estimated as

approximately 10,800 L (over 144 min), which is less than that in section ①.

As shown in Figure 7, Rayleigh frequency shifts with time are obtained via a 1A cable and

FIMT fiber. The maximum frequency shift was near −7.5 GHz, which is equal to +50 με strain,

thereby indicating that the formation deformation was in an expansive state. A significant

notable leap occurred at 204.4–205.0 m and was attributed to the packers effect.

As discussed in section ②, Figure 8 shows the geomechanical deformation profile of the

formation accompanying the water injection. Rayleigh frequency shifts were sampled every

17.4 min employing the 1A cable and plotted as a 2D contour. Vertical extension of the

impacted zone focused on an area of 202.2-209.2 m in accordance with the strain response

©2019 American Geophysical Union. All rights reserved.


range shown in the red dashed line in Figure 7. The underlying formation was wider than the

injection zone (up to ~2.8 m). Thus, when the injection rate varied from 60 L/min to 90

L/min, a higher strain occurred, and thereby increased the impacted zone. Furthermore, an

obvious signal hiatus at 204.9 m was observed in the deformation profile, which is consistent

with the strain trough shown in Figure 7. This was interpreted as the effect of the coupling

protector as previously mentioned regarding section ②. Thus, cable installation technology

plays a vital role in obtaining high-quality data for DFOS-based field monitoring at large-scale

application sites.

4.3 Numerical Simulation of Water Injection Processes

To estimate strain responses resulting from two water injections, coupled thermal–

hydraulic-mechanical (THM) models, multiphase flow (TOUGH2) and geomechanical

(FLAC3D), were combined in a series. The coupled simulator TOUGH2-FLAC3D is used to

analyze fluid flow accompanied by deformation in geothermal studies and seismicity as

triggered by natural fluid pressure (Rutqvist et al., 2002). Based on the geological profile

shown in Figure 2, a simplified model containing two water injections is established as

shown in Figure 9. The model was set up on the basis of an axisymmetric horizontal layered

model with a geometric size of 300 × 300 × 300 m, and the temperature of the upper

boundary is corresponded to 10 ℃ with a geothermal gradient of 3 ℃/100 m. Normal

displacement for the other boundaries is set to zero in thermal isolation and with no flow,

and the major model parameters are listed in Table 2. Practically, the numerical model in the

study constitutes part of the model used in Lei et al., (2019). For sections ① and ②,

©2019 American Geophysical Union. All rights reserved.


injection scenarios were uniformly distributed amongst the grid cells corresponding to the

sandy layers adjacent to the injection well. Injection rates and pressures were set based on

the observed data shown in Figures 3 and 6.

Simulated impacted zones near injection formations and the corresponding curves of strain

history at three time points are shown in Figure 9 (𝜀𝑧𝑧 denotes the vertical strain). It is

drawn that induced expansive deformation (i.e., positive strain) occurred in the injection

formations, and the impacted zones spread into the surrounding formations beyond the

injection sections, thereby implying that water injection inevitably decreased effective

pressure, expansive strain occurrence, and ultimately ground uplift. With the exception of a

slight deviation in strain magnitude of the actual observed data, the simulated vertical strain

profiles were generally in accordance with the measured strain response trends (see Figures

4 and 7). Thus, the simulation results can also further provide insights on the effectiveness

of this hybrid DFOS technology for in-situ monitoring of small geomechanical deformations,

i.e., strain variations in subsurface water injection. However, it should be noted that the

hydraulic characterization of injection layers are not analyzed in detail in the study, because

Lei et al., (2019) analytically and numerically demonstrated that sensitive and scaled rock

properties including equivalent permeability and pore compressibility can be adequately

constrained by water head and distributed strain data recorded via the DFOS system during a

pumping test performed at the same site in the study. Although testing types (pumping

water vs injection water) are different in different ranges (in addition to two wells with 5 m

apart used in the study, two pumping wells with 175 m and 280 m apart with MW #1 in Lei

©2019 American Geophysical Union. All rights reserved.


et al., 2019), the essences of geomechanical deformation are the same. Thus, fluid

production or injection alters the stress field of injected formations and causes strain

changes, which are sensed effectively with the DFOS technology. This proves the

applicability of DFOS for monitoring reservoir formation responses at different near-wellbore

regions in a manner consistent with previous field testing work (Lei et al., 2019; Xue et al.,

2018). Furthermore, in conjunction with the time-dependent resistivity logs of MW #1 and

IW #2, it is deduced that small-scale heterogeneity and discontinuities in the formations can

lead to strain changes and vertical expansion of the impacted zones. However, it is extremely

difficult to quantify and assign small-scale formation heterogeneity, such as shale-sand ratio,

to each grid element in detail.

5 Discussions

Based on induced geomechanical deformation profiles acquired by the DFOS cable in two

water injection tests, the two main factors that were analyzed included formation conditions

associated with the injection sections and impacted zones, and injection scenarios. Given

the changes in the injection scenarios, the results indicated that the injection rate and

pressure agree well with the magnitudes and states of induced strains and ranges of the

impacted zones as reported in extant studies (Aamri et al., 2017; Heffer, 2002).

Approximately, if the upper formation corresponds to a clay- dominated lithology (i.e.,

caprock), then the geomechanical responses during fluid injection can interpret the

potential for caprock fracturing and permanently assess the wellbore integrity and caprock

sealing in depth (Zoback & Gorelick, 2012).

©2019 American Geophysical Union. All rights reserved.


5.1 Explanation of Geomechanical Deformation Responses

5.1.1 Effect of Lithologic Heterogeneity

To better understand the effect of lithologic heterogeneity on the geomechanical

deformation (i.e., strain) induced by water injection, many geophysical well logging

measurements were simultaneously performed at the open-hole, production casing

insertion, and fluid injection stages, particularly gamma-ray logging in the borehole.

Given that radioactive isotopes are typically associated with the clay minerals in shales, it is a

commonly accepted practice to use the relative gamma-ray deflection as a shale volume

indicator. Gamma-ray index 𝐼GR is defined as a linear scaling of the 𝐺𝑅 between 𝐺𝑅max

and 𝐺𝑅min as follows:

𝐺𝑅−𝐺𝑅min
𝐼GR = (6)
𝐺𝑅max −𝐺𝑅min

where 𝐺𝑅, 𝐺𝑅max and 𝐺𝑅min denote the actual gamma value in the zone of interest (API),

in the shale zone (𝐼GR = 1), and in the clean zone-no shale (𝐼GR = 0), respectively.

The shale content is calculated using eq. (6). However, the results exhibit significant error

and should be recalibrated. The alternative relationship is expressed in terms of 𝑉sh as

follows:

2𝐻 ∙𝐼GR −1
𝑉sh = (7)
2𝐻 −1

where 𝑉sh denotes the shale content (%) and H is a constant with its value typically set to 2

for Mesozoic and older rocks (Larionov, 1969).

©2019 American Geophysical Union. All rights reserved.


The diagram of 𝑉sh -GR is shown in Figure 10. Generally, a siltstone formation exhibits a

higher 𝑉sh value than that of sandstone, and thus the shale volume along the borehole can

be estimated using open-hole GR data. Hence, to better understand the effect of lithologic

heterogeneity on the geomechanical deformation (i.e., induced strain) state of a

reservoir-caprock system being considered for CO2 storage in a region, many geophysical

logging tools with applications to wellbore integrity are implemented in the borehole. As

shown in Figure 10, the impacted regions over time expand in both vertical directions up to

2.8 m. When compared with the borehole histogram (BH) in the impacted regions, it is

observed that the shale content significantly affects the magnitude of the induced strain and

deformation expansion. Decreases in 𝑉sh related to the formation allows for easier strain

accumulation, thereby indicating that injected CO2 is more prone to storage and migration in

sand-rich rocks. In the first injection of section ②, a thin sand interlayer (approximately

184.5–184.7 m) exists that allows for strain when the water injection commences. Similarly,

the siltstone (189.2–189.8 m) below the injection zone is unable to block the strain

expansion. Thus, although the impact acting on clay-rich formations (i.e., higher 𝑉sh ) was

relatively small, the existence of the siltstone only weakened the injection-induced

deformation intensity as opposed to completely stopping the spread of the induced strain.

5.1.2 Effect of Injection Scenario

From Figure 10, it can be evidently speculated that the impacted zones extended toward the

vertical direction during water injection with the exception of the corresponding injection

section in MW #1. The main reason is that water injection can lead to stress arching effects

©2019 American Geophysical Union. All rights reserved.


wherein the vertical stress becomes heterogeneous across the reservoir. Given poroelastic

deformation, reservoir dilatancy during injection can increase vertical strain and bulk

modulus change within the reservoir. This can potentially be transmitted to the upper and

lower zones, or even to the surface to be manifested as ground uplift (Chen, 2011; Shirzaei

et al., 2016). Another factor corresponds to the sand thickness, which is an important factor

that affects the magnitude and spreading extent of induced strain. For example, maximum

strains of 87 με at ① and 70 με at ②. Moreover, within the 5.5 m distance between MW #1

and IW #2, injected water migrates along the 5.5 m-long sand layer and gradually pushes the

original formation water to the fibered well with different flow patterns as shown in Figure

11. Therefore, the pressure front of the injected water propagates in the formation although

the induced deformation logged by the DFOS exhibits different responses under the two

injection scenarios. At section ①, the water front quickly spreads to the fibered well such

that the deformation profile exhibits a consistent response along the depth. At section ②,

the water front slowly moves to the fibered well over time, and the interface of first-arrival

water pressure front is portrayed in the profile as denoted in cyan in Figure 10 (Read et al.,

2014). Notably, the different strain responses from 184.5 to 184.6 m and 205.0 to 205.7 m

are derived from the clamps’ and coupling protectors’ effects, respectively, as previously

stated in section 4. Concurrently, itis easy to infer that increases in injection volume (Q = Vit)

and injection pressure induced expansive strain in and near the injection formations.

Therefore, in order to quantitatively characterize the spatial-temporal distribution of

geomechanical deformation in subsurface during the water injection, time-lapse strain

©2019 American Geophysical Union. All rights reserved.


curves at seven monitoring points of 1A cable are shown in Figure 12 to depict the temporal

evolution of induced strain at some certain depths. Strain histories of seven observed points

(Figure 12) and injection volume changes exhibit increase in strain magnitudes with injection

although there were some signs of gradual decline after water injection ceased. Second,

strain fluctuation is generally consistent with the pressure increment, which can

approximately estimate fluid injectivity. Third, in comparison with strain data in two

sections, it is observed that geomechanical deformation is vertically distribute more

uniformly at section ① with propagation from 187.8 m to 189.9 m (overlying point) and

191.1 m (underlying point) (Figure 12a) when compared to that at section ② with maximum

value of 207.3 m (underlying point) (Figure 12b) mainly due to a coupling protector

mentioned in Section 4.2, although section ① exhibited higher injection volume and larger

pressure variations as a whole. Fourth, at the given depth in Figure 12, the maximum strain


curves corresponded to a green line with a peak value of 𝜀t,max = 65.9 με at 187.8 m (Figure


12a) and yellow line with a peak value of 𝜀t,max = 50 με at 207.3 m (Figure 12b). At the given

time in Figures 4 and 7, the maximum strain profiles corresponded to both cyan lines with

① ② ① ①
peak values of 𝜀d,max = 60 με and 𝜀d,max = 43 με, respectively. Hence, [𝜀V,max ⁄𝜀t,max ](d=187.8

① ①
m) = 55.6 με/65.9 με = 0.84 [𝜀V,max ⁄𝜀d,max ](t=228.3 min) = 55.6 με/60 με = 0.93 as well as

① ① ① ①
[𝜀V,max ⁄𝜀t,max ](d=207.3 m) = 42.2 με/50 με = 0.843[𝜀V,max ⁄𝜀d,max ](t=141.3 min) = 42.2 με/43 με =

0.98, indicating that strain proportion potentially varied with short temporal and spatial

scales.

©2019 American Geophysical Union. All rights reserved.


To further interpret different vertical extension of the impacted zones in the fibered well,

the open-hole resistivity logging data of two wells are shown in Figure 13. Given a dip angle

of ~10° between adjacent wells, the injection formations can extend downwards, as shown

in the blue dashed lines in Figure 13. Thus, this can potentially explain the downward

extension of the induced geomechanical deformation of the impacted zones beyond the

corresponding injection sections. Alternatively, formation resistivity around the injection

sections rarely changes, as further evidenced by the relative shale content of 10–25% within

the range. This means that once the water is injected, it inevitably migrates outside the

injection formations, and thus results in vertical extension of the impacted zones when

water flows to MW #1. With the exception of effects from different injection pressures, the

maximum impacted zones actually hinged on the formation differences. From the resistivity

logging curve, the average resistivity of ② is lower than that of ①, thereby indicating that

section ② exhibits less shale and higher permeability and allows more fluid to inject with a

lower deformation fluctuation.

Alternatively, in the impacted zone, the strain changes with time exceeded that of the

injection section. It is speculated that the extension of sand formation via the injecting water

can affect the upper and lower siltstone formations. Hence, it is necessary to further validate

the results in the future by conducting numerical simulations. More interestingly, when the

water injection started and stopped, the corresponding strains also concurrently emerged

and disappeared. In addition, the impacted zone in the second test is larger than the first

test, thus proving that the induced range depends on injection volume in addition to

©2019 American Geophysical Union. All rights reserved.


lithologic heterogeneities.

5.2 Remarks on GCS Field Monitoring

In a manner similar to water injection, the core component of GCS involves injecting

supercritical CO2 (scCO2) into deep reservoir formations to mitigate the global greenhouse

effect. As widely known, CO2 injection also alters the geomechanical field accompanying the

temperature, strain, or pressure redistribution. Therefore, over the GCS life cycle, it will be of

great significance to in-situ monitor the signs to track fluid behaviors and potential

consequences from the surface to the subsurface such as ground uplift at GCS sites.

Evidently, novel application of a proven DFOS is an optimal benchmark to map subsurface

deformation and the scCO2 footprint in an analogy to the water injection. Links between

surface deformation and injected subsurface reservoirs will be established from a

perspective of continuous in-situ strain responses using only one fiber optic cable. Hence,

the investigation can significantly resolve geomechanical modeling deficiencies associated

with CO2 injection, such as simplified models that ignore effects of lithologic heterogeneity

on surface deformation based on the poroelastic mechanism of reservoir–caprock systems.

The strain information reveals implications for the nature of formation deformation and

caprock integrity induced by fluid injection near a wellbore. This provides many

opportunities for wider applications in geomechanics and hydrology.

6 Conclusions

©2019 American Geophysical Union. All rights reserved.


In the study, a distributed optical fiber sensing system integrating Brillouin and Rayleigh

backscattering techniques was utilized to cross-well monitor water injection tests. The

first-hand results from the DFOS can be used to account for the magnitude and range of

geomechanical deformation along the entire well depth direction. Thus, the study can

indicate significant implications for interwell monitoring at CGS-related sites via DFOS

technology. Based on the methodology, a few conclusions are obtained as follows:

• An advanced DFOS system based on hybrid Brillouin–Rayleigh backscattering was

successfully applied for real-time monitoring of water injection tests between two

adjacent wells. Two different injection formations ① and ② were investigated.

• Field testing demonstrated that the induced geomechanical deformation (i.e., strain) was

mainly dependent on the injection scenario with the exception of the injection

temperature, which did not significantly affect strain changes.

• Geomechanical deformation concurrently emerges and disappears with injection scheme,

although the effect of packers on the strain profile results in the presence of small strain

leaps. The impacted zones expand along the vertical direction beyond the corresponding

injection sections. Furthermore, the injection-induced strains are positive, showing that

deformation of the impacted zones behaves expansive deformation throughout all the

tests. Hence, the DFOS approach visualizes the deformation of the targeted formation and

the extent of deformation under different injection scenarios.

• Formation heterogeneity results in lateral transport of injected water and its pressure

propagation in the pore spaces (Baatz et al., 2015; Rutqvist, 2012). When compared with

©2019 American Geophysical Union. All rights reserved.


𝑉sh -GR and BH, the lower 𝑉sh related to the formation allows for easier strain

accumulation. Additionally, although the impact on the clay-rich formations is relatively

small, the existence of shale only weakens the induced deformation intensity as opposed

to completely stopping strain expansion.

• Based on the open-hole resistivity logs of the IW and MW, the simplified simulation result

further confirms the strong small-scale formation heterogeneity and discontinuities of the

shale–sand distribution in the injection formations.

• Testing results indicate that there are two main reasons for the vertical expansion of

impacted zones during the water injections. The first is the formation heterogeneity,

which is quantified by the open-hole resistivity logging and GR data and mainly affects

fluid behaviors including migration, evolution, and diffusion in the subsurface. The second

reason is poroelastic deformation given that different injection scenarios can generate

different injection pressure variation patterns, which ultimately lead to different

geomechanical deformation responses and first-arrival pressure front patterns of the

injected water propagating to the fibered well.

It was identified that this hybrid DFOS system, in the case of deep subsurface monitoring,

can be deployed well behind the casing for real-time in-depth in-situ monitoring of fluid

behavior for applications in future carbon capture and storage (CCS) and enhanced

geothermal system (EGS) projects with the aim to address wellbore failure, CO2 leakage, and

hydraulic stimulation in or near the observation wells. Additionally, these investigations can

also guide the field layout of fiber cables and data interpretation in wellbores, which

©2019 American Geophysical Union. All rights reserved.


constitute in-depth knowledge reserved for in-situ monitoring applications of CO2 storage

sites.

Overall, the study mainly centers on characterizing the induced strain profiles in real time

and in situ during the water injection using the proposed hybrid DFOS system and especially

in elaborate interpretations for unequal vertical strain expansions. However, the

hydromechanical properties of injected sediments significantly constrain pressure response

and fluid transfer. Thus, future studies can aim to quantify their contributions in detail in

terms of dominating downhole deformation measurements of fiber cables, although two

recently published articles discussed in this aspect in detail (Lei et al., 2019; Zhang et al.,

2019). Furthermore, it is essential to point out that no other independent measurements of

induced strain responses at injection sites that can be used to directly compare with DFOS

results, and this will necessitate installing cable-matching tools to integrate

continuously-acquired underground data with monitored information such as bottom hole

pressure, seismic and micro-seismic. Therefore, further refinement for cable

instrumentation should also be investigated in future work to open massive opportunities to

be explored in hard-to-access locations, such as deep subsurface energy storage regions.

Acknowledgments

This paper is based on results obtained from a project commissioned by the New Energy and

Industrial Technology Development Organization (NEDO) and the Ministry of Economy,

Trade and Industry (METI) of Japan. We would like to thank staff of NBX and RITE involved in

this project. All the data used in this study are publicly available at

©2019 American Geophysical Union. All rights reserved.


https://github.com/SunRITE20170913/Dataset2019WR024794R. Correspondence and

requests for materials should be addressed to Yankun Sun. We are also grateful for the

constructive comments and recommendations from the Associate Editor, Professor Tetsu

Tokunaga, and four anonymous reviewers, who helped us to substantially improve our

manuscript.

References

Aamri, M., Al Zadjali, R., Mahajan, S., & Hindriks, C. (2017). Geomechanical Assesment of Injection Pressure

Constraints for a Waterflood Field Development. SPE Reservoir Characterisation and Simulation

Conference and Exhibition. Abu Dhabi, UAE: SPE. https://doi.org/10.2118/185969-MS

Arnon, A., Lensky, N. G., & Selker, J. S. (2014). High-resolution temperature sensing in the Dead Sea using fiber

optics. Water Resources Research, 51(2012), 341–358. https://doi.org/10.1002/2013WR014979

Baatz, R., Bogena, H. R., Franssen Hendricks, H.-J., Huisman, J. A., Montzka, C., & Vereecken, H. (2015).

Geomechanics of subsurface water withdrawal and injection. Water Resources Research, (51), 5974–

5997. https://doi.org/10.1002/2014WR015608

Bachmann, C. E., Wiemer, S., Goertz-Allmann, B. P., & Woessner, J. (2012). Influence of pore-pressure on the

event-size distribution of induced earthquakes. Geophysical Research Letters, 39(9), 1–7.

https://doi.org/10.1029/2012GL051480

Bakku, S., Wills, P., & Fehler, M. (2014). Monitoring Hydraulic Fracturing Using Distributed Acoustic Sensing in

a Treatment Well. Proceedings of the Society of Exploration Geophysicisits Annual Meeting, Denver,

Colorado, 26-31 October, 5003–5008. https://doi.org/10.1190/segam2014-1559.1

Baldwin, C. S. (2014). Brief history of fiber optic sensing in the oil field industry. In Fiber Optic Sensors and

©2019 American Geophysical Union. All rights reserved.


Applications XI (Vol. 9098, pp. 909803–909809). SPIE. https://doi.org/10.1117/12.2072360

Bissell, R. C., Vasco, D. W., Atbi, M., Hamdani, M., Okwelegbe, M., & Goldwater, M. H. (2011). A full field

simulation of the in Salah gas production and CO2 storage project using a coupled geo-mechanical and

thermal fluid flow simulator. Energy Procedia, 4, 3290–3297.

https://doi.org/10.1016/j.egypro.2011.02.249

Cappa, F., Guglielmi, Y., Gaffet, S., Lançon, H., & Lamarque, I. (2006). Use of in situ fiber optic sensors to

characterize highly heterogeneous elastic displacement fields in fractured rocks. International Journal of

Rock Mechanics and Mining Sciences, 43(4), 647–654. https://doi.org/10.1016/J.IJRMMS.2005.09.016

Cappa, Frédéric, Guglielmi, Y., Rutqvist, J., Tsang, C. F., & Thoraval, A. (2008). Estimation of fracture flow

parameters through numerical analysis of hydromechanical pressure pulses. Water Resources Research,

44(11), 1–15. https://doi.org/10.1029/2008WR007015

Cappa, Frédéric, Scuderi, M. M., Collettini, C., Guglielmi, Y., & Avouac, J.-P. (2019). Stabilization of fault slip by

fluid injection in the laboratory and in situ. Science Advances, 5(3), eaau4065.

https://doi.org/10.1126/sciadv.aau4065

Chen, Z. R. (2011). Poroelastic model for induced stresses and deformations in hydrocarbon and geothermal

reservoirs. Journal of Petroleum Science and Engineering, 80(1), 41–52.

https://doi.org/10.1016/j.petrol.2011.10.004

Dakin, J. P., Pratt, D. J., Bibby, G. W., & Ross, J. N. (1985). Distributed optical fibre Raman temperature sensor

using a semiconductor light source and detector. Electronics Letters, 21(13), 569–570.

https://doi.org/10.1049/el:19850402

Delepine-Lesoille, S., Guzik, A., Bertrand, J., Henault, J.-M., & Kishida, K. (2013). Validation of TW-COTDR

©2019 American Geophysical Union. All rights reserved.


method for 25km distributed optical fiber sensing. In Fifth European Workshop on Optical Fibre Sensors

(Vol. 8794, p. 879438). SPIE. https://doi.org/10.1117/12.2025802

Duguid, A., Guo, B., & Nygaard, R. (2017). Well Integrity Assessment of Monitoring Wells at an Active CO2-EOR

Flood. Energy Procedia, 114(November 2016), 5118–5138.

https://doi.org/10.1016/j.egypro.2017.03.1667

Freifeld, B., Daley, T., Cook, P., Trautz, R., & Dodds, K. (2014). The modular borehole monitoring program: A

research program to optimize well-based monitoring for geologic carbon sequestration. Energy Procedia,

63, 3500–3515. https://doi.org/10.1016/j.egypro.2014.11.379

Fukuda, K., Suzuki, M., & Ito, M. (2015). The origin and internal structures of submarine-slide deposits in a

lower Pleistocene outer-fan succession in the Kazusa Forearc Basin On The Boso Peninsula of Japan.

Sedimentary Geology, 321, 70–85. https://doi.org/10.1016/j.sedgeo.2015.03.009

Guglielmi, Y., Cappa, F., Lançon, H., Janowczyk, J. B., Rutqvist, J., Tsang, C. F., & Wang, J. S. Y. (2015). ISRM

Suggested Method for Step-Rate Injection Method for Fracture In-Situ Properties (SIMFIP): Using a

3-Components Borehole Deformation Sensor BT 2007-2014. In R. Ulusay (Ed.), Rock Mechanics and Rock

Engineering (pp. 179–186). Cham: Springer International Publishing.

https://doi.org/10.1007/978-3-319-07713-0_14

Hartog, A. H. (2017). An introduction to distributed optical fibre sensors. An Introduction to Distributed Optical

Fibre Sensors. https://doi.org/10.1201/9781315119014

Heffer, K. (2002). Geomechanical influences in water injection projects: An overview. Oil and Gas Science and

Technology, 57(5), 415–422. https://doi.org/10.2516/ogst:2002027

Kersey, A. D. (2000). Optical fiber sensors for permanent downwell monitoring applications in the oil and gas

©2019 American Geophysical Union. All rights reserved.


industry. IEICE Transactions on Electronics, 83(3), 400–404.

Kishida, K., Li, C. H., Nishiguchi, K., Yamauchi, Y., Guzik, A., & Tsuda, T. (2012). Hybrid Brillouin-Rayleigh

distributed sensing system. In SPIE (Vol. 8421, pp. 84212G-8421–4). https://doi.org/10.1117/12.975668

Kishida, K., Yamauchi, Y., & Guzik, A. (2014). Study of optical fibers strain-temperature sensitivities using hybrid

Brillouin-Rayleigh system. Photonic Sensors, 4(1), 1–11. https://doi.org/10.1007/s13320-013-0136-1

Kogure, T., & Okuda, Y. (2018). Monitoring the Vertical Distribution of Rainfall-Induced Strain Changes in a

Landslide Measured by Distributed Fiber Optic Sensing With Rayleigh Backscattering. Geophysical

Research Letters, 45(9), 4033–4040. https://doi.org/10.1029/2018GL077607

Kurashima, T., Horiguchi, T., & Tateda, M. (1990). Distributed-temperature sensing using stimulated Brillouin

scattering in optical silica fibers. Optics Letters, 15(18), 1038–1040.

https://doi.org/10.1364/OL.15.001038

Larionov, V. V. (1969). Radiometry of Boreholes. NEDRA. Moscow: NEDRA.

Lei, X., Xue, Z., & Hashimoto, T. (2019). Fiber Optic Sensing for Geomechanical Monitoring: (2)- Distributed

Strain Measurements at a Pumping Test and Geomechanical Modeling of Deformation of Reservoir

Rocks. Applied Sciences, 9(3), 417. https://doi.org/10.3390/app9030417

Moore, J. R., Gischig, V., Button, E., & Loew, S. (2010). Rockslide deformation monitoring with fiber optic strain

sensors. Natural Hazards and Earth System Science, 10(2), 191–201.

https://doi.org/10.5194/nhess-10-191-2010

Read, T., Bour, O., Selker, J. S., Bense, V. F., Borgne, T. Le, Hochreutener, R., & Lavenant, N. (2014).

Active-Distributed Temperature Sensing to continuously quantify vertical flow in boreholes. Water

Resources Research, (50), 3706–3713. https://doi.org/10.1002/2014WR015273

©2019 American Geophysical Union. All rights reserved.


Ringrose, P., Atbi, M., Mason, D., Espinassous, M., Myhrer, \O, Myhrer, Ø., et al. (2009). Plume development

around well KB-502 at the In Salah CO2 storage site. First Break, 27(7005), 85–89.

Rutqvist, J., Wu, Y.-S., Tsang, C.-F., & Bodvarsson, G. (2002). A modeling approach for analysis of coupled

multiphase fluid flow, heat transfer, and deformation in fractured porous rock. International Journal of

Rock Mechanics and Mining Sciences, 39(4), 429–442. https://doi.org/10.1016/S1365-1609(02)00022-9

Rutqvist, Jonny. (2012). The geomechanics of CO2 storage in deep sedimentary formations. Geotechnical and

Geological Engineering, 30(3), 525–551. https://doi.org/10.1007/s10706-011-9491-0

Rutqvist, Jonny, Cappa, F., Rinaldi, A. P., & Godano, M. (2014). Modeling of induced seismicity and ground

vibrations associated with geologic CO2 storage, and assessing their effects on surface structures and

human perception. International Journal of Greenhouse Gas Control, 24, 64–77.

https://doi.org/10.1016/j.ijggc.2014.02.017

Sanada, H., Sugita, Y., & Kashiwai, Y. (2012). Development of a multi-interval displacement sensor using Fiber

Bragg Grating technology. International Journal of Rock Mechanics and Mining Sciences, 54, 27–36.

https://doi.org/10.1016/j.ijrmms.2012.05.020

Schenato, L. (2017). A Review of Distributed Fibre Optic Sensors for Geo-Hydrological Applications. Applied

Sciences (Vol. 7). https://doi.org/10.3390/app7090896

Shanafield, M., Banks, E. W., Arkwright, J. W., & Hausner, M. B. (2018). Fiber-optic Sensing for Environmental

Applications: Where We’ve Come From- and What’s Possible? Water Resources Research.

https://doi.org/10.1029/2018WR022768

Shirzaei, M., Ellsworth, W. L., & Tiampo, K. F. (2016). Surface uplift and time-dependent seismic hazard due to

fluid injection in eastern Texas. Science, 2122(2010), 2118–2122.

©2019 American Geophysical Union. All rights reserved.


https://doi.org/10.1126/science.aag0262

Sun, Y., Li, Q., Yang, D., Fan, C., & Sun, A. (2016). Investigation of the dynamic strain responses of sandstone

using multichannel fiber-optic sensor arrays. Engineering Geology, 213.

https://doi.org/10.1016/j.enggeo.2016.08.008

Sun, Yankun, Li, Q., Fan, C., Yang, D., Li, X., & Sun, A. (2017). Fiber-optic monitoring of evaporation-induced

axial strain of sandstone under ambient laboratory conditions. Environmental Earth Sciences, 76(10), 379.

https://doi.org/10.1007/s12665-017-6706-6

Sun, Yankun, Li, Q., & Fan, C. (2017). Laboratory core flooding experiments in reservoir sandstone under

different sequestration pressures using multichannel fiber Bragg grating sensor arrays. International

Journal of Greenhouse Gas Control, 60, 186–198.

https://doi.org/https://doi.org/10.1016/j.ijggc.2017.03.015

Sun, Yankun, Xue, Z., & Hashimoto, T. (2018). Fiber optic distributed sensing technology for real-time

monitoring water jet tests: Implications for wellbore integrity diagnostics. Journal of Natural Gas Science

and Engineering, 58, 241–250. https://doi.org/10.1016/j.jngse.2018.08.005

Tyler, S. W., Selker, J. S., Hausner, M. B., Hatch, C. E., Torgersen, T., Thodal, C. E., & Schladow, S. G. (2010).

Environmental temperature sensing using Raman spectra DTS fiber-optic methods. Water Resources

Research, 46(4), 1–11. https://doi.org/10.1029/2008WR007052

Verdon, J. P., Kendall, J.-M., Stork, A. L., Chadwick, R. A., White, D. J., & Bissell, R. C. (2013). Comparison of

geomechanical deformation induced by megatonne-scale CO2 storage at Sleipner, Weyburn, and In

Salah. Proceedings of the National Academy of Sciences , 110(30), E2762–E2771.

https://doi.org/10.1073/pnas.1302156110

©2019 American Geophysical Union. All rights reserved.


White, J. A., Chiaramonte, L., Ezzedine, S., Foxall, W., Hao, Y., Ramirez, A., & McNab, W. (2014). Geomechanical

behavior of the reservoir and caprock system at the In Salah CO2 storage project. Proceedings of the

National Academy of Sciences , 111(24), 8747–8752. https://doi.org/10.1073/pnas.1316465111

Xue, Z., & Hashimoto, T. (2017). Geomechanical Monitoring of Caprock and Wellbore Integrity Using Fiber

Optic Cable: Strain Measurement from the Fluid Injection and Extraction Field Tests. Energy Procedia,

114, 3305–3311. https://doi.org/10.1016/j.egypro.2017.03.1462

Xue, Z., Park, H., Kiyama, T., Hashimoto, T., Nishizawa, O., & Kogure, T. (2014). Effects of hydrostatic pressure

on strain measurement with distributed optical fiber sensing system. Energy Procedia, 63, 4003–4009.

https://doi.org/10.1016/j.egypro.2014.11.430

Xue, Z., Shi, J.-Q., Yamauchi, Y., & Durucan, S. (2018). Fiber Optic Sensing for Geomechanical Monitoring:

(1)-Distributed Strain Measurements of Two Sandstones under Hydrostatic Confining and Pore Pressure

Conditions. Applied Sciences, 8(11), 2103. https://doi.org/10.3390/app8112103

Yoshiaki Yamauchi. (2010). A measurement method to determine strain and temperature coefficients in fiber

optic sensors. In Proc. of APWSHM Conference, Tokyo, Japan. Tokyo.

Zhang, Y., Xue, Z., Park, H., Shi, J. Q., Kiyama, T., Lei, X., et al. (2019). Tracking CO2 plumes in clay-rich rock by

distributed fiber optic strain sensing (DFOSS): a laboratory demonstration. Water Resources Research,

(55), 856–867. https://doi.org/10.1029/2018WR023415

Zoback, M. D., & Gorelick, S. M. (2012). Earthquake triggering and large-scale geologic storage of carbon

dioxide. Proceedings of the National Academy of Sciences , 109(26), 10164–10168.

https://doi.org/10.1073/pnas.120247310

©2019 American Geophysical Union. All rights reserved.


Table 1. Configuration properties of the water injection tests in IW #2.

Measured scatteringa Cable categoryb


ID Injection section Lithology
R B 3A 3T 1A C

① 186.8~188.8 m Sand layer with thin mud layer √ √ √ √ ×

② 204.4~206.4 m Sand layer √ √ √ × √ ×

a
R: Rayleigh measurement; B: Brillouin measurement.

b
3A: 3 layer armored cable; 3T: temperature fiber (FIMT); 1A: single layer armored cable; C: cable used for

communication only.

©2019 American Geophysical Union. All rights reserved.


Table 2. Major hydraulic and mechanical properties used in the model.

Property Bulk Shear Pore Sand Silt Formation

modulus modulus compressibility permeability permeability porosity

Unit GPa GPa Pa-1 mD mD ---

Value 0.80 0.40 3.9 x 10-9 2500 10 0.43

©2019 American Geophysical Union. All rights reserved.


(a) N (b)

5.5 m
Japan MW #1 IW #2

Mobara

0 200 m

(c)

10 m

(d)
Water injection lead

IW #2
P-T sensor
MW #1

Figure 1. Description of the test sits in this study. (a) Location of the water injection test site in Mobara,

Japan. (b) Two wells (fiber installed monitoring well MW #1 and water injection well IW #2) configuration

at the site marked in yellow dashed line and red circles, respectively. (c) 5.5 m apart between MW #1 and

IW #2. (d) Double-packer for the water injection.

©2019 American Geophysical Union. All rights reserved.


IW #2 MW #1
Injection well Monitoring well
Fiber behind casing
MD
0m

Perforation interval 5.5 m


149.4 m
151.4 m
Fiber cable

188.8 m Casing
Cement
204.4 m
149 m

206.4 m Water
P/T sensor
209.4 m Water

211.8 m

213.8 m
230 m

230.0 m

Sand
` 300 m
Silts

Figure 2. Schematic illustration of the water injection tests between two adjacent wells. MD: Measured

Depth, the same below.

©2019 American Geophysical Union. All rights reserved.


Figure 3. The injection rate and pressure with injection time at section ①. The triangles in different colors

indicate the injection time points corresponding to the water injection time of the strain responses shown

in Figure 4.

©2019 American Geophysical Union. All rights reserved.


185 m

191 m

Figure 4. Strain responses in depth under different injection time at section ①. Significant impacted zone

is 185~ 191 m with initial strain of 5 με in red dashed line. Note that these strain values here were just

calculated using Rayleigh frequency shift, similarly hereinafter.

©2019 American Geophysical Union. All rights reserved.


Figure 5. Induced geomechanical deformation profiles of the injection formations at section ①. In Track 1,

EMI image shows a laminated sand/shale sequence of the formations in MW #1, generally brighter yellow

colors for more resistive sand-rich facies while darker brown colors for less resistive shale–rich facies.

Track 2 indicates the formation resistivity changes. In Track 3, the borehole histogram (BH) is plotted

based on EMI and resistivity logs. In Tracks 4 and 5 display the deformation profiles measured by DFOS

system at 170~220 m and 183~195 m, respectively. (For interpretation of the references to color in this

figure legend, the reader is referred to the web version of this article.)

©2019 American Geophysical Union. All rights reserved.


Figure 6. The injection rate and pressure with injection time at section ②. The triangles in different colors

indicate the injection time points corresponding to the water injection time of the strain responses shown

in Figure 7.

©2019 American Geophysical Union. All rights reserved.


202.2 m

209.2 m

Figure 7. Strain responses in depth under different injection time at section ②. Significant impacted zone

is 202.2~ 209.2 m with initial strain of 5 με in red dashed line.

©2019 American Geophysical Union. All rights reserved.


Figure 8. Induced geomechanical deformation profiles of the injection formations at section n. Track 1:

EMI. Track 2: Resistivity logs. Track 3: Borehole histogram. Tracks 4 and 5: Deformation profiles measured

by DFOS system at 170~220 m and 201~212 m, respectively. (For interpretation of the references to

color in this figure legend, the reader is referred to the web version of this article.)

©2019 American Geophysical Union. All rights reserved.


Figure 9. Numerical simulation of water injection tests at sections ① (a) and ② (b). I. Simplified

simulation models for injection sections ① (upper) and ( (lower). II. Vertical strain profiles induced water

injection. III. Plots present for induced strain histories at injection time of 5 min, 50 min and 200 min,

respectively. In the models, injection sections (green), sand formations (blue) and silt formations (red) are

roughly modeled. In the model, average permeability of injection formation ①, ② and the silt layer set

2.5 D, 3.0 D and 10.0 mD, respectively.

©2019 American Geophysical Union. All rights reserved.


-87 με 87 με
184
185
1.8 m
186

Injection zone
186.8~188.8
187

m
188
189
190
191 2.2 m

-70 με 70 με
202
203 2.2 m
·

Injection zone
204.4~206.4
204
205

m
206
207
208
209 2.8 m
210

BH

Figure 10. Open-hole gamma-ray log (red) and calculated shale volume (blue) at the 170-220 m depth of

the fibered well (MW#1, OF-1, 300 m in Mobara, Japan) with comparison with the borehole histogram

(BH). The geomechanical deformations of the injection formations at sections ① and ② are plotted on

the right.

©2019 American Geophysical Union. All rights reserved.


Fiber cable Casing

Cement

Cement
4
DFOS interrogator

Pressure fronts
Water injection well Monitoring well

Mud layer
Mud layer
Water ` Water

Water Water

Pressure fronts
Mud layer Fiber cable

Sand layer

Figure 11. Conceptual diagram of potential water seepage pathways and its pressure front patterns

between two adjacent wells.

©2019 American Geophysical Union. All rights reserved.


Figure 12. Time-lapse strain response histories recorded at different depths as a function of injection

volume (Q) and pressure increment (ΔP =Pt-P0) during the water injection at sections ① (a) and ② (b). a,

Strain histories in depths of upper layers at 184.7 m (yellow) and 185.9 m (dark yellow), injection layers at

186.8 m (blue), 187.8m (green) and 188.8 m (magenta), and lower layers at 189.8 m (cyan), 191.1 m (red),

respectively. b, Strain histories in depths of upper layers at 202.6 m (cyan) and 203.5 m (magenta),

injection layers at 204.4 m (blue), 205.6 m (red) and 206.4 m (green), and lower layers at 207.3 m (yellow)

and 208.4 m (dark yellow), respectively. Note that the time points denoted with dash lines are consistent

with those of depth-dependent strain profiles (Figures 4 and 7) in the same colors and their

corresponding pressures marked near the dash lines.

©2019 American Geophysical Union. All rights reserved.


IW #2 MW #1

Extended formation
Injection

Impacted
Impacted zone
Injection
Extended

Figure 13. Open-hole resistivity logs of MW #1 (red) and IW #1 (blue). The resistivity changes in IW #2,

MW #1 corresponding to the injection formations and maximum impacted zones are denoted in dashed

rectangles in cyan, magenta and yellow, respectively. The possible impacted zones and the extended

formations in MW #1 corresponding to injection sections ① and ② are shown in cyan and magenta

dashed lines.

©2019 American Geophysical Union. All rights reserved.

You might also like