You are on page 1of 13

Journal of Hydrology 624 (2023) 129981

Contents lists available at ScienceDirect

Journal of Hydrology
journal homepage: www.elsevier.com/locate/jhydrol

Research papers

Particle blockage mechanism of a weakly consolidated sandstone


geothermal reservoir during tailwater recharge
Jieqin Xia a, b, c, Bin Dou a, b, c, *, Hong Tian a, b, c, Peng Xiao a, b, c, Jun Zheng a, b, c, Xiaotian Lai a, b, c
a
Faculty of Engineering, China University of Geosciences, Wuhan 430074, China
b
National Center for International Research on Deep Earth Drilling and Resource Development, Wuhan 430074, China
c
Key Laboratory of Deep Geothermal Resources, MNR, China University of Geosciences, Wuhan 430074, China

A R T I C L E I N F O A B S T R A C T

Keywords: Weakly consolidated sandstone geothermal reservoirs (WCSGR) are widely distributed, highly developed and
Weakly consolidated sandstone geothermal utilized, with tremendous potential to meet enormous energy needs. However, this kind of geothermal reservoir
reservoirs (GR) suffer from permeability damage during tailwater recharge, primarily due to particle blockage (PB). The
Tailwater recharge
exact impact of PB on the permeability of GRs during tailwater recharge remains unclear. Therefore, an
Particle blockage
experimental study investigating PB in WCSGR tailwater recharge was conducted, considering the comprehen­
Permeability
Blockage mechanism sive effects of multiple factors. The study aimed to understand the mechanism of PB and its effect on GR
permeability by observing the migration of fluorescent particles in sandstone core. Through analysis of micro­
sphere migration/blockage properties, theoretical calculations, and empirical criteria, the PB mechanism during
tailwater recharge in weakly consolidated sandstone was revealed. A physical model of particle movement was
developed to explain the permeability attenuation phenomena in the experimental cores, describing the
occurrence and development process of PB in real recharge operations. Finally, the theoretical computation were
integrated with the experimental findings, and the chemical and dynamic conditions of geothermal tailwater
recharge in the WCSGR were optimized. This study provides theoretical guidance for further numerical simu­
lations and field tests.

1. Introduction (Oliveira et al., 2014). At the same time, the hydraulic gradient gener­
ated during operation can cause the dislodge of some reservoir particles
Hydrothermal geothermal resources offer advantages such as easily, which are then transported deeper into the WCSGR and ulti­
shallow burial depth, low exploitation difficulty, wide applicability, and mately redeposited. On the other hand, the recharging fluid disrupts the
significant potential for development and utilization. The weakly balance of the thermal, hydrodynamic, stress, and chemical fields in the
consolidated sandstone geothermal reservoir (WCSGR), as one of the original reservoir, which causes particle migration and swelling,
typical hydrothermal resource reservoirs (Anderson and Rezaie, 2019), compaction, scaling, and even blockage in the reservoir. Particle
is characterized by abundant reserves, and enormous development po­ blockage (PB) greatly reduces the recharge efficiency in WCSGR projects
tential potential (Kong et al., 2014). A WCSGR generally refers to a and indirectly affects the utilization benefits of geothermal reservoir
loosely porous sandstone reservoir with a uniaxial compressive strength (GR) energy (Han et al., 2020). PB forms rapidly and has the greatest
(UCS) between 5 MPa and 10 Mpa (Cui et al., 2022; Heiland and Flor, impact on the permeability damage in WCSGR (Bouwer, 2002; Caselles-
2006; Zhang et al., 2022b; Yan et al., 2019). It can be approximated as a Osorio et al., 2007; Goss et al., 1973; Platzer and Mauch, 1997; Zhang
porous system of a solid rock skeleton, spatial pore structure, and liquid- et al., 2022a). It is the primary factor hindering the srecharge of the
phase fluid, with relatively loose sand grains. Geothermal recharge WCSGR. Therefore, the PB mechanism during WCSGR recharge is a very
technology is the key to ensuring the sustainable development of important subject of research, and the development of targeted, sus­
WCSGR resources (Kamila et al., 2021). In the process of WCSGR tainable, and breakthrough optimization measures for blockage relief is
recharge, on the one hand, drilling and completion operations can alter crucial to overcoming the technical bottleneck of WCSGR recharge.
the physical and chemical properties of the rocks in the near-well area Various factors affecting particle migration and deposition have been

* Corresponding author at: Faculty of Engineering, China University of Geosciences, Wuhan 430074, China.
E-mail address: doubin@cug.edu.cn (B. Dou).

https://doi.org/10.1016/j.jhydrol.2023.129981
Received 18 March 2023; Received in revised form 7 June 2023; Accepted 17 July 2023
Available online 21 July 2023
0022-1694/© 2023 Elsevier B.V. All rights reserved.
J. Xia et al. Journal of Hydrology 624 (2023) 129981

researched, i.e., ionic strength (Bennacer et al., 2017; Torkzaban et al.,


2015), particle size (Hu and Brusseau, 1994), seepage velocity (Liu et al.,
2016), pore structure (Ahfir et al., 2009; Ahfir et al., 2017; Bai et al.,
2016), hydrodynamic forces (Chen and Bai, 2015), and gravity (Ahfir
et al., 2017). In aaddition, Lun (2006) found that in core flow experi­
ments, pore throats are easily but slowly blocked when the particle
concentration is low and the ratio of pore size to particle size (pore size
ratio) is greater than 3. The pore size ratio and particle concentration
determine the initial and final blockages, respectively. Rehg et al. (2005)
observed a correlation between particle size and particle deposition
depth, suggesting that larger particles,are more likely to deposit on the
surface of the infiltrated medium. Zhao et al. (2019) further studied the
migration characteristics of particles (3.24 μm diameter) in water-
saturated medium with different concentrations (100 mg/L, 300 mg/
L, 500 mg/L) by performing sand column recharge experiments. They
found that surface blockage is formed through the combined action of
filtration and precipitation, while internal blockage is controlled by
filtration, precipitation, inertia effect, and hydraulic effect. Further­
more, Chai et al. (2022) recently studied the characteristics and mech­
anisms of formation damage in a sandstone GR during decreased salinity Fig. 1. Average volume density curve of the rock samples granularity.
water injection through multiple experiments. Additionally, the flow
rate (Badalyan, 2015; Vigneswaran and Suazo, 1987; You et al., 2019), was equivalent to that of the fluorescent microspheres so that the mi­
temperature (Bai et al., 2017; Rosenbrand et al., 2015; Rosenbrand and crospheres were uniformly dispersed in the brine medium.
Fabricius, 2014; Wang et al., 2021a; Wang et al., 2021b), salt concen­
tration (Chai et al., 2022; Chequer et al., 2018; Khilar and Fogler, 1984; dcs = (0.15 − 0.41)Dcs (1)
Kia et al., 1987; Liu et al., 2020), and pH (Ochi and Vernoux, 1998b;
Yang et al., 2022) of the recharging fluid have been found to affect PB. 2.2. Experimental schemes
Although studies on PB have been abundant, most experimental
studies on particle transport and blockage during sandstone GR recharge The high-pressure and high-temperature core flow instrument and
have focused on the influence of a single factor, neglecting the mutual the rock samples were used to simulate the dynamic response process of
coupling effects among all the factors. To gain a better insight into the WCSGR tailwater recharge. The instrument consists of six units, as
causes of PB, we conducted dynamic simulations of the WCSGR tail­ depicted in Fig. 2. Prior to the actual seepage experiment, five steps were
water recharge process under the integrated effect of multiple factors followed: assembling the rock sample, removing air, applying confining
using a high-pressure and high-temperature core flow experimental in­ pressure, removing the vacuum, and saturating the rock with water.
strument. By employing polystyrene fluorescent microspheres and an Polystyrene fluorescent microspheres and an ultraviolet analyzer
ultraviolet analyzer, we visualized particle transport, and explored the (Fig. 3) were used to visualize particle transport. The experimental
characteristics of PB during GR tailwater recharge. Finally, according to scheme, as presented in Table 1, and the five variables of flow rate,
the empirical criterion before the experiment, the phenomenon after the confining pressure, microsphere size, turbidity, and temperature were
experiment, and the theoretical analysis, the PB mechanism and devel­ considered. These variables aimed to explore the variation characteris­
opment process during tailwater reinjection in the WCSGR were clari­ tics of permeability in sandstone recharge under the combined effect of
fied. Prevention and control measures that are conducive to the on-site multiple factors. The experimental sampling area had a recharge flow
recharge project were proposed. rate of 20–30 m3/h, with a recharge depth of 660 m and a recharge well
diameter of 216 mm. Consequently, the experimental flow rate was
2. Materials and methods approximately 0.46 mL/min, while 0.5 mL/min was assumed in this
experimental scheme. It’s worth noting that seepage experiments were
2.1. Materials conducted after draining the airusing a vacuum pump, neglecting the
influence of temperature and pressure changes that may release bubbles
The experimental rock samples used in the study were weakly during the experiment.
consolidated sandstone obtained from the Guantao Formation in Lan­
kao. These rock samples had dimensions of φ25mm × 50 mm. The depth 2.3. Empirical method of judging blockage
from which they were collected ranged approximately
1891.73–1899.47 m. The particle size distribution of the samples is 2.3.1. Evaluate the risk of particle blockage
shown in Fig. 1. The average volume density means the percentage of The membrane filtration index (MFI), proposed by Dillon (Dillon
the volume content of the particles of this size in the total particles. The et al., 2001b), was used to initially evaluate the blockage risk of pore
average pore throat size (dcs ) among the coarse particles (number of throats in the experimental rock by fluorescent microspheres. MFI refers
valleys between the second and third peaks) was calculated by empirical to the rate at which suspended particles are blocked when passing
formula (1) (Zhang et al., 2020), where Dcs is the average diameter of through a diameter of 0.45 μm. A higher MFI index indicates a greater
coarse particles and 0.15 and 0.41 are empirical values. Logging data risk of blockage caused by fluorescent microspheres, and the quantita­
suggest that the chemical type of geothermal water is sodium chloride tive relationship is as follows:
(Cl− -Na+). Therefore, the experimental water samples consisted of a
MFI = 15.2 × TSS (2)
high salinity NaCl solution with a concentration of 8 wt% (weight
percent). Fluorescent microspheres were also included in the solution. where MFI is the blocking potential index of fluorescent micro­
The high salinitysolution could inhibit the swelling of clay minerals, i.e., spheres, s/L2; TSS is the turbidity of fluorescent microspheres, mg/L;
montmorillonite and illite-montmorillonite mixed layers in the pores of and 15.2 is the empirical value.
rock samples and ensure that particle separation and migration are only There is no unified standard for discerning the blockage potential
caused by fluid flow. At the same time, the density of the injected fluid with MFI values (Dillon et al., 2001a). Van Duijvenbode and Olsthoorn

2
J. Xia et al. Journal of Hydrology 624 (2023) 129981

Fig. 2. Core flow experimental instrument.

Fig. 3. Experimental materials and instruments.

presented MFI values not exceeding 20 s/L2 in the Amsterdam water 2.3.2. Diagnostic model of particle blockage
supply, while the Netherlands specified a range of 3–5 s/L2 for The PB mode can be divided into four categories: surface filter cake
rechargeable fine sand media. Dillon et al. (2001b) concluded that a risk blockage (S), internal pore throat blockage (P), double blockage (S-P),
of blockage arises when MFI values reach at least 110 s/L2. However, in and no blockage (N). The quantitative discriminant relationship among
the experimental design of fluorescent microspheres with turbidities of the porous medium, particle size, and blocking mode can be expressed as
2, 4, 10, 20, and 30 mg/L, and the corresponding MFI values were (Glover and Walker, 2009):
calculated as 30.4, 60.8, 152, 304, and 456 s/L2. Therefore, even at very √̅̅̅̅̅̅̅̅̅
low turbidity levels, the risk of blockage cannot be negligible. am2 DP′
Deff = 2θ • Reff = 2⋅ ⋅ (3)
8n2m 2

3
J. Xia et al. Journal of Hydrology 624 (2023) 129981

Table 1 permeability)/initial permeability, and the evaluation index of the


Laboratory experimental schemes of dynamic recharge. blockage degree is shown in Table 3 (Hu, 2012).
Experiment Flow Confining Microsphere Turbidity Temperature
group rate pressure size (µm) (mg/L) (℃) 3. Results and analysis
(mL/ (Mpa)
min)
3.1. Phenomenon after the experiments
A B C D E

1 0.5 0.8 3 2 25 3.1.1. Dynamic change characteristics of permeability


2 0.5 1 5 4 35 The permeability damage rate after the experiments is in the range of
3 0.5 1.2 10 10 45 33.33%-77.05%. According to Table 4, the damage degree was deter­
4 0.5 1.4 20 20 55
5 0.5 1.6 30 30 65
mined to range from moderately weak to strong. In addition, it was
6 1 0.8 5 10 55 found that the fluorescent microsphere turbidity of 2 mg/L still causes
7 1 1 10 20 65 damage to the core permeability, which further confirms the results of
8 1 1.2 20 30 25 the blockage risk evaluation.However, the degree of damage is different
9 1 1.4 30 2 35
under different experimental conditions. Fig. 3 shows the dynamic
10 1 1.6 3 4 45
11 2 0.8 10 30 35 change curves of the core permeability of groups (1), 7, 13, 19, and 25
12 2 1 20 2 45 during horizontal seepage lasting 4 h. The values taken for each factor in
13 2 1.2 30 4 55 the experiment were different, reflecting the change in core perme­
14 2 1.4 3 10 65 ability under the influence of comprehensive factors. The corresponding
15 2 1.6 5 20 25
core permeability damage rates for these groups are 57.14%, 50.45%,
16 3 0.8 20 4 65
17 3 1 30 10 25 50.00%, 37.74%, and 33.82%, respectively. The degree of blockage is
18 3 1.2 3 20 35 moderately strong for the first three groups and moderately weak for the
19 3 1.4 5 30 45 latter two groups. In the experiments, the core permeability of the five
20 3 1.6 10 2 55
groups decreases to different degrees increasing seepage time. The
21 4 0.8 30 20 45
22 4 1 3 30 55
decline rate decreases gradually and finally tends to be stable, resem­
23 4 1.2 5 2 65 bling an “exponential function”. In actual tailwater recharge, fine par­
24 4 1.4 10 4 25 ticles continuously migrate deeper into the formation and eventually
25 4 1.6 20 10 35 deposit in a stable location within the reservoir, leading to a progressive
decrease in reservoir permeability during continuous recharge opera­
where Deff is the particle size of the particles constituting the porous tion. In contrast, during the experimental process, the experimental core
is short, the process from particle transport blockage to blockage
structure, mm, which is taken as D50 ; θ is the dimensionless conversion
removal is relatively balanced, and the core permeability gradually
factor; n is the porosity, dimensionless; m is the cementation index be­
becomes stable with increasing seepage time, without showing a mul­
tween the media particles, which is taken as 1.5, dimensionless; DP′ is the
tiplicative change.
effective pore diameter, mm; DP is the particle size of the injected par­
In Fig. 4, Further local observations reveal that the permeability
ticles, mm; and the value of a is 8/3. Then, according to the judgment
variation curves exhibit fluctuations, which can be attributed to the
criteria ofDP′/Dp ≤ 1 for S, 1 less than 180DDPp < for S-P, andDP′/Dp ≥ 180

complex transport behaviors of unstable particles in the core pore
for P. Then, the PB mode here is S-P, and the detailed data are shown in throats within a continuous kinetic environment. Particularly in group
Table 2. (7), there is a noticeable position in the curve where the local amplitude
experiences a sudden increase and then gradually stabilizes. This is a
2.4. Evaluation method of blockage degree typical fluctuation caused by the behavior of particle transport blockage
and unblocking. In the cases of group (13), the position of the sudden
The flow state of the fluid in the granular materials’ pores can be rise in the curve is due to a partial reduction in the applied confining
determined by the Reynolds number. In the above 25 groups of exper­ pressure on the core during the experiment. The release of pressure
iments, the Reynolds numbers are in the range of 0.14–2.72, indicating makes the partial recovery of elastic deformation in the pore throat,
the Darcy’s law could be used to calculate the fluid flow. The absolute resulting in a rapaid increase in permeability, which then gradually
permeability was selected as the method to measure the permeability. decreases to reach a stable stage.
In Fig. 5, the experimental data of groups 7 and 25 conform to a
KΔPA double exponential decay function, and the curve fitting effect is good.
Q= (4)
μL The decay rate of permeability in the first 50 min of the two groups’
where μ is the viscosity of the fluid, mPa⋅s; A is the cross-sectional experiment is significantly faster than that in the last 200 min.
area of a porous medium, cm2; L is the length of the porous medium, Furthermore, the decay rate of permeability in the first 25 min of group
cm; ΔP is the pressure difference in a laminar state, MPa; and Q is the 25 is faster than that in group (7). This difference could be attributed to
flow rate per unit time through the rock pore, mL/min. the larger size of fluorescent microspheres in the injected fluid of group
The permeability of the rock samples can be inversely calculated by 25 and the smaller core permeability. Under the influence of a high flow
Formula 4, where A, ΔP, L, μ, and Q are known quantities in the rate, the size filtration effect of the fluorescent microspheres causes
experiment. The experimental results are expressed as the permeability rapid blocking of the core surface, commonly known as the “filter cake
damage rate (%), which is equal to (initial permeability-final effect”. This phenomenon not only demonstrates that S blockage is the
blockage mode with the fastest reduction in recharge efficiency during
Table 2
geothermal recharge but also confirms that the relative size difference
Discrimination of particle blocking mode. between the injected particles and the pore throats has a significant
influence on the permeability decay rate. Moreover, the data dispersion
Particle size of fluorescent microspheres 3 5 10 20 30
(µm)
states of sections A and B in the two experimental groups are different,
and the decreasing rate of section A in group 25 is significantly larger
D′P /Dp 12.28 7.37 3.68 1.84 1.23
than that in group (7). This can be interpreted that group (7) has a larger
Blocking mode S-P
pore throat size, smaller injected microsphere particle size, and lower

4
J. Xia et al. Journal of Hydrology 624 (2023) 129981

Table 3
Evaluation index of blockage degree.
Permeability damage rate(%) ≤0.05 0.06–0.30 0.31–0.50 0.51–0.70 0.71–0.90 >0.90

Blockage degree None Weak Moderately weak Medium to strong Strong Extremely strong

microspheres with a size of 30 μm were deposited at the bottom of the


Table 4
intermediate vessel. This observation aligns with the findings reported
Related parameters for the calculation of FE .
in the literature (Song et al., 2021; Song et al., 2020), where it was found
Parameter Value Parameters Value that particles with a median size of 24.29 μm tend to accumulate at the
Hamaker constant, 5.2×10− 21
Relative dielectric 2.56 bottoms of suspension tanks, pipelines, and inlet chambers of sand
N•m 43.875 constant of solution 8.854×10− 12 tanks. The deposition of particles at the bottom can be explained by the
Ionic strength in − 26 Vacuum dielectric 0.5×10− 9 DLVO theory (Muneer et al., 2020). According to this theory, when the
solution, mol/L − 50 constant 1.6×10− 19
Surface potential, mV Atomic collision
concentration of colloidal particles is high, the repulsive barrier of the
Pore wall potential, diameter, m potential energy curve decreases, and only an obvious primary potential
mV Elementary electric well appears, which means that the colloidal particles are highly un­
charge stable and more susceptible to sedimentation or deposition.
The experimental combinations with similar core permeabilities and
the same turbidity were selected for comparison and analysis, i.e., group
(14) and group 17 (turbidity of 10 mg/L), group (15) and group 18
(turbidity of 20 mg/L), and group (11) and group 22 (turbidity of 30
mg/L). The retention of fluorescent microspheres in the samples after
the experiment was observed by UV analyzer, and the results are shown
in Figs. 6-8. From a macroscale perspective, it was observed that the
fluorescent microspheres did not penetrate the samples and mostly
accumulated near the entrance of the samples. Several factors can
explain this phenomenon. Firstly, the hydrophobicity of fluorescent
microspheres is very strong in high-salinity solutions, and the internal
structure of rock samples is complex, making it difficult for the micro­
spheres to easily penetrate the samples. The fact that the microspheres
were observed to move through the relatively smooth surface walls of
the samples and the rubber tube further supports this explanation.
Additionally, the pressure difference between the front and back of the
samples may be too small for the microspheres to travel deeper into the
samples (Bizmark et al., 2020). It is also possible that the concentration
of fluorescent microspheres flowing out with the fluid is too low to be
easily observed with the naked eye. Morover, the longitudinal com­
parison among Figs. 6-8 shows that turbidity has the most obvious effect
on the distribution of the fluorescent microspheres. Higher the turbidity
Fig. 4. Experimental permeability changes of groups (1), 7, 13, 19 and 25.
leads to easier gathering of microspheres in the surface layer of the
samples. The horizontal comparison indicates that higher flow rate at
fluid flow rate compared to those of group 25. As a result, the effect of
the same turbidity can promote the transport of fluorescent micro­
unstable particle transport deposition within a short time on the overall
spheres into the interior of the samples, thereby slowing the surface
permeability of the core is weaker in group (7) compared to group 25.
aggregation density. However, it is important to note higher turbidity
(particle concentration) can result in the formation of severe PB on the
3.1.2. Transport characteristics of fluorescent microspheres
surface of porous media, even if the flow rate increases. According to the
During the experiment under a low flow rate (group (5), it was
retention of fluorescent microspheres in the samples, the blocking
observed that some highly concentrated polystyrene fluorescent

Fig. 5. Fitting curves of experimental groups 7 and 25.

5
J. Xia et al. Journal of Hydrology 624 (2023) 129981

Fig. 6. The distribution of fluorescent microspheres in groups 14 and 17 (turbidity of 10 mg/L) after the experiment.

Fig. 7. The distribution of fluorescent microspheres in groups 15 and 18 (turbidity of 20 mg/L) after the experiment.

Fig. 8. The distribution of fluorescent microspheres in groups 11 and 22 (turbidity of 30 mg/L) after the experiment.

Fig. 9. Morphology and microstructure of rock samples before the experiment. Note: The reflective areas in Fig. 9 are the gold powder sprinkled on the samples
before scanning.

6
J. Xia et al. Journal of Hydrology 624 (2023) 129981

modes are judged to be S and S-P, which has a certain deviation from the are variable and uncontrolled under the effect of synthetic factors, so the
theoretical diagnosis results (all S-P blocking mode) in section 2. This forms of particle migration and blockage are very diverse.
demonstrates that the pore size ratio of the samples can only prelimi­
narily determine the blocking mode, and it is also closely related to the 3.2. Theoretical analysis
particle concentration and flow rate.
In order to investigate the retention of fluorescent microspheres in 3.2.1. Physicochemical effects
the rock sample, scanning electron microscopy (SEM) was performed on In the process of particle recharge with fluid, the motion of particles
the samples before and after the experiment, as shown in Figs. 9-11. is jointly determined by both physicochemical and hydrodynamic ef­
Taking group 17 as an example, Fig. 10 intuitively shows the internal fects (Ochi and Vernoux, 1998a). These effects are influenced by factors
structure of the sandstone after the tailwater reinjection experiment. By such as particle size, dispersion distance, and other factors. The physi­
compared with the microscopic scanning results before the experiment cochemical effects include the Brownian motion of colloidal particles
(Fig. 9) and after the experiment (Fig. 10), while disregarding the themselves and the adsorption phenomenon due to the interaction of
interference of exogenous fluorescent microspheres, it can be observed particles when they are close to the surface of sand (Kim and Benefield,
that the overall internal structure of the experimental samples became 1996). Brownian motion plays an important role in the migration of
slightly disordered, with more clay particles or movable sand grains particles smaller than 1 μm, as their movement is strongly infuenced by
around the rocks (Fig. 10). This indicates that some endogenous parti­ thermal agitation. However, for particles larger than 1 μm, the effect of
cles within the sample have been dispersed, detached, and transported brownian motion is not obvious due to the higher viscosity of the fluid.
under the influence of the fluid. There are complicated interaction The mutual repulsion or attraction between particles smaller than 10 μm
mechanism between particles, between particles and pore walls, and and sand grains depends on comprehensive forces (Muneer et al., 2020),
between particles and fluid. Thesemechanisms compete with each other, i.e., van der Waals attraction (long-range force, FL ), electric double-layer
resulting in particles eventually manifesting in three different forms: repulsion (FD ), Born repulsion force (short-range force, FB ), and acid-
random dispersion on the rock surface, aggregation in pore throat base interaction (FAB ). These forces can be calculated with Eqs. (5) to
channels, or penetration through the rock sample with the fluid flow. (10) (Cui et al., 2022; Khilar and Fogler, 1998).
Furthermore, as shown in Fig. 10, pore throat blockage is caused by two [ ( )]
types of particles, endogenous movable sand grains and exogenous A132 a a D′
FL = − + + ln (5)
fluorescent microspheres. The former can directly block the entire pore 6 D′ D′ + 2a D′ + 2a
throat channel, which is extremely harmful to the internal pores of the (
samples. The latter mainly has three manifestation forms: adsorption, FD =
bridge blocking, and screening-filtering. The numbers 1–5 in Fig. 10 [ ( ) ])
mark the fluorescent microspheres adsorbed on the surface of the me­ 1 + exp( − κh) ( )
∈ a/4 2ψ 1 ψ 2 ln + ψ 21 + ψ 22 ln(1 − exp( − 2κh)
dium. At present, fluorescent microspheres do not directly block the 1 − exp( − κh)
pore throat, but they pose a potential risk. Continuous recharge may (6)
transport the fluorescent microspheres and subsequently lead to pore [ ]
throat blockage. Regions around positions A and B are further enlarged A132 (σ)6 8 + H 6− H
FB = + (7)
in Fig. 10, revealing the “bridge blocking” phenomenon of the fluores­ 7560 a (2 + H)7 H7
cent microspheres at position A, and screening-filtering occurring at the
pores due to size effects at position B. The formation process and action [ (
h − h0
)]
mechanism of the injected fluorescent microspheres are described in
0
FAB = ∓FAB exp − (8)
λ
detail in section 4. In addition, Fig. 11 demonstrates that some micro­
( )
spheres adsorbed on the surface of the movable sand grains were kB T
transported along with them to the pore throats and remained there. ψ = ζ = − 2.3 (PH − PH 0 ) (9)
e
Additionally, some microspheres carried slightly dispersed kaolinite
minerals adsorbed on the surface of the rocks. In summary, the internal 2e2 NA I 1/2
structure and mineral composition of a WCSGR are very complex, and κ=( ) (10)
∊kB T
the transport trajectories of both endogenous and exogenous particles
where a is the radius of spherical particles, m; D and D’ are the surface

Fig. 10. Internal microscopic view of the rock sample after the experiment (magnification 300 times on the left and 1000 times on the right).

7
J. Xia et al. Journal of Hydrology 624 (2023) 129981

Fig. 11. Retention of fluorescent microspheres (particle size 30 µm) (magnification 1000 times).

separation distance of equal-diameter particles and the distance from


the spherical particles to the plane, m; A132 is the Hamaker constant; h is
the dispersion distance, m; and ∊ is the dielectric constant, J− 1C2m− 1. ψ
is the surface potential, and its value is not easily measured; thus, it is
generally replaced by the zeta potential. The value of ζ can be directly
measured or can be used for reference by the Nernst formula (9). T is the
absolute temperature, K; e is the number of charges, e = 1.60 × 10− 19 C;
PH0 is the equipotential point of the surface; and NA is the Avogadro
constant, NA = 6.02 × 1023 . I is the ionic strength of the solution, which
is related to the solution concentration Ci and the ionic valence Zi , that

is, I = 0.5 Ci Z2i . σ is the atomic collision diameter at the Lennard-
Jones potential, σ = 0.5 nm; H = h/a, which is much less than 1.
Meanwhile, when h>σ, FB drops sharply and can be assumed to be
0
negligible relative to FD and FL . FAB is the hydrophobic interaction, J,
and its value is approximately tens to hundreds of kB T, but the highly
hydrophobic solid may reach 2900 kB T. h0 is the minimum equilibrium
distance, m; λ is the attenuation length of liquid molecules, m. Taking
water as an example, its h0 and λ are 0.5 nm and 1.0 nm, respectively.
Calculation and Analysis
The theoretical calculation is based on appropriate simplifications:
the curvature of the pore surface is neglected, the particle is regarded as Fig. 13. FE calculation results of 3 µm microspheres.
a standard sphere, and the pore surface is considered a plane–plate
structure. Considering that the short-range force and the long-range
As shown in Fig. 13, the values of FD , FB , and FAB tend to be zero after
force act on the particle surface within approximately 5 nm and 100
a distance of 4 nm, and the particles are only subject to the action of FL .
nm, respectively, and any hydrated ions in the solution can offset the
This is because the high concentration of the seepage solution, which
natural repulsion, which makes the particle contact distance of
leads to a thin electric double layer. Thus, the repulsion potential barrier
approximately 0.3 nm (Zamani and Maini, 2009).Therefore, the mini­
in the FE curve completely disappears, and the gravitational potential
mum distance h between the particle and the pore surface is taken as 0.3
energy dominates. Consequently, the particles are easily adsorbed on the
nm. Experimental group (10) (pH = 6.5, T = 45℃) is taken as an
surface of the rock medium, which is consistent with the experimental
example, as it has characteristics similar to geothermal recharge water
phenomenon described in section 3.1.2. Furthermore, the FD values of 3
in the Lankao area. It is reasonable to assume a sphere–plate geometry
μm microspheres at different temperatures (Fig. 14) indicate that the
(Fig. 12), and FE (the sum of FL , FD , FB andFAB ) is calculated for 3 μm
higher the temperature is, the larger the FD , leading to greater damage to
particles at different distances h. The related parameters for the calcu­
the core permeability.These calculated results are in agreement with the
lation of FE are shown in Table 4.
experimental results and previous studies (Cui et al., 2019).

3.2.2. Hydrodynamic effects


Hydrodynamic effects describe how the viscosity of the fluid affects
the particle migration rate. The flow velocity in sand pores varies from
maximum at the center to zero near the surface, leading to the exertion
of shear forces on the particles by the flowing water. thus further
influencing the transport properties of the particles. Generally, the hy­
drodynamic force is a volumetric force and plays a significant role in the
separation and transport of particles larger than 10 μm. It includes
gravity (G), buoyancy (Fw ), the drag force (Fd ), friction, the lift force (Fs ),
the false mass force (Fm ), the pressure gradient force (Fp ), and the
Magnus force (FMa ). These forces can be calculated based on equations
(11)-(16) (Goldman et al., 1967; O’Neill, 1968; Tripathy, 2010).
G = ρP Vp g (11)
Fig. 12. Simplified calculation system.

8
J. Xia et al. Journal of Hydrology 624 (2023) 129981

microspheres may be trapped on the surface of rock samples due to the


size filtering effect or migrate to somewhere inside the core with the
action of fluid, and then form a blockage. These findings align with the
SEM results described in section 3.1.2 after the completion of the
experiment.

3.2.3. Calculation and analysis of comprehensive effects


The actual situation of 3, 5, 10, and 20 μm microspheres in the
process of fluid transport are further clarified based on the above
calculation methods, and the results are shown in Table 6. Comparing
the data from Tables 5 and 6, it can be observed that the resultant forces
acting on 3 μm and 5 μm microspheres, which are adsorbed on the
surface of sand grains, are negative in the vertical direction. In contrast,
the resultant forces on 10 μm, 20 μm, and 30 μm microspheres are
positive, indicating their separation from the sand grains. It is worth
noting that the Fv of experimental group (3) in Table 6 is negative. This
could be due to the low flow velocity and density in the seepage envi­
ronment, where the buoyancy and lifting forces are insufficient to
separate the microspheres from the sand grains. Alternatively, the Fs
calculated by the formula might be smaller than the actual experimental
Fig. 14. FD calculation results of 3 µm microspheres at different temperatures. value. In summary, the experimental results suggest that 3 μm and 5 μm
microspheres are easily adsorbed on the surface of sand grains, which
Fw = ρf Vp g (12) directly reduces the pore throat area. However, 10 μm, 20 μm, and 30
μm microspheres are prone to migration with the action of hydrody­
Fd = 6 × 1.7009πμav1 − vp (13) namics, eventually causing blockages in the pore throats. These pro­
cesses collectively contribute to the decrease in permeability within
√̅̅̅̅̅̅̅̅̅̅
⃒ ⃒̅ sandstone reservoirs.
√̅̅̅̅̅̅̅̅ 2 ⃒⃒ ⃒ ⃒duf ⃒
FS = 1.61 ρf μ dp uf − up ⃒⃒ ⃒⃒
⃒ (14) In addition, it should be noted that the high salt concentration of the
dy
recharging fluid used in the experiment eliminates the swelling of clay
minerals in the process of recharge and ensures that the density of the
∂p 8 μQ seepage solution closely matches that of the fluorescent microspheres.
Fp = − V P = − VP 4 (15)
∂l πr0 However, this high salt concentration actually promotes the adsorption
of reservoir particles and injected microspheres onto the surface of sand
1
FMa = πdp3 ρf ωuf − up (16) grains. It is necessary to clarify the real movement state of particles in
8
the process of GR recharge and exclude the influence of this high
where ρP and ρf are the particle density and solution density, respec­ salinity. Therefore, the 3 g/L mineralization degree of the GR ground­
tively, kg/m3; Vp is the particle volume, m3; g is the acceleration of water is converted to a molar concentration of 0.05 mol/L. The forces
gravity, g = 9.8m/s2 ; μ is the fluid viscosity, Pa⋅s; a is the particle radius, acting on the microspheres are recalculated, and the results are listed in
m; v1 and vp are the real flow velocity of fluid (equal to flow rate/ Table 7. It can be observed that the adsorption forces on the 3 μm and 5
porosity) and particle transport velocity, m/s, respectively; ω is the
μm microspheres decrease, while the repulsive forces on the 10 μm, 20
rotational angular velocity of particles ; VP is the volume of particles, uf
μm, and 30 μm microspheres increase. This demonstrates that the mi­
crospheres are more likely to migrate under low salt concentrations.
and up are the instantaneous velocities of fluid and particles, respec­
However, the movement behavior of microspheres, as described above
tively; r0 is the radius of the rock sample, cm; and Q is the inflow rate per
without considering the effect of salt concentration, remains consistent.
second.
Therefore, the theoretical calculation results can still be used to reveal
Calculation and Analysis
the PB mechanism during tailwater reinjection in a WCSGR.
The sand grains are simplified as standard spheres, fluorescent mi­
crospheres are assumed to be frictionless due to their smooth surface.
4. Discussion
Additionally, some forces that have little effect on particles are ignored
(Fig. 12(2)). The force Fv (N) and torque M (N⋅m) acting on 30 μm mi­
4.1. Mechanism of particle blockage in tailwater reinjection
crospheres during fluid flow at pH 6.5 (solution potential ζ of − 26 mV)
are then calculated. FE is taken as the gravitational maximum, and it is at
Based on the dcs calculated in section 2.1 and the empirical criterion
the primary potential well of the total interaction potential energy. The
of “1/2–1/5” (Yu et al., 2018), the particle penetration through the rock
results listed in Table 5 indicate that the hydrodynamic effects (G + Fw
samples can be preliminarily determined as follows. The particle sizes of
+ Fs + Fd + Fp) exert a greater influence on 30 μm microspheres
the 3 μm and 5 μm microspheres are smaller than 1/5 of the minimum
compared to the physicochemical effects (FE). This is manifested as the
dcs of the samples, and thus, these microspheres are in the state of free
mutual separation between particles and sand grains. Therefore, the

Table 5
The calculation results of forces acting on 30 μm microspheres.
13 13
Group number G(10− ) Fw (10− ) Fs (10− 8) FE (10− 9) Fd (10− 10
) Fp (10− 4) Fv (10− 8) M(10− 8)

Group 17 1.4824 1.5028 4.3711 − 3.1993 26.4062 6.0766 4.0512 1.8229


Group 9 1.4986 1.3082 − 3.2381 7.1153 0.5458 0.9940 1.6371
Group 21 1.4933 4.7492 − 3.1906 23.5281 7.2191 4.4402 2.1656
Group 13 1.4871 2.1739 − 3.0393 9.9004 1.5189 1.8700 0.4556
Group 5 1.4799 5.0102 − 2.9879 2.1137 0.0811 0.2022 0.0243

9
J. Xia et al. Journal of Hydrology 624 (2023) 129981

Table 6
The calculation results of Fvertical on each microsphere in the experiment.
10 10 10 10
3 µm Fvertical (10− ) 5 µm Fvertical (10− ) 10 µm Fvertical (10− ) 20 µm Fvertical (10− )

Group 1 − 2.5791 Group 15 − 5.4215 Group 24 0.5403 Group 8 0.4340


Group 18 − 3.2381 Group 2 − 5.3226 Group 11 0.1853 Group 25 0.2116
Group10 − 3.1906 Group 19 − 5.2403 Group 3 − 3.7735 Group 12 0.8490
Group 22 − 3.1330 Group 6 − 5.1476 Group 20 0.2604 Group 4 3.8618
Group 14 − 3.0825 Group 23 − 5.0638 Group 7 1.1118 Group 16 0.0114

Table 7
The calculation results of Fvertical on each microsphere in the geothermal reservoir.
11 11
3 µm Fvertical (10− ) 5 µm Fvertical (10− ) 10 µm Fvertical (10− 9) 20 µm Fvertical (10− 8) 30 µm Fvertical (10− 8)

Group 1 − 4.8685 Group 15 − 8.2034 Group 24 6.3080 Group 8 0.6136 Group 17 4.3199
Group 18 − 4.7110 Group 2 − 8.0583 Group11 2.7423 Group 25 2.2923 Group 9 1.2579
Group 10 − 4.6194 Group 19 − 7.9054 Group 3 0.4979 Group12) 1.0226 Group 21 4.6999
Group 22 − 4.5466 Group 6 − 7.7827 Group 20 3.4639 Group 4 0.2093 Group 13 2.1252
Group 14 − 4.4667 Group 23 − 7.6479 Group 7 0.9569 Group16 1.3043 Group 5 0.4532

migration; the particle sizes of 10 μm, 20 μm, and 30 μm microspheres separation and migration of particles is predominantly affected by
fall between 1/2 and 1/5 of the minimum dcs , and these microspheres colloidal and hydrodynamic forces. Colloidal forces (McCarthy and
show a migration–deposition state. However, the empirical formula Zachara, 1989) are essentially electrostatic force and can be further
mentioned above considers only the relationship between particle size divided into van der Waals attraction, electric double layer repulsion,
and pore throat size and neglects the influence of physicochemical and ect. On the other hand, hydrodynamic forces (Ochi and Vernoux, 1998a)
hydrodynamic effects on particle behavior that will change their are related to the flow characteristics of the recharging fluid. The force
movement state during actual recharge operations (Frey et al., 1999; analysis of particles is depicted in Fig. 16. Here, FE represents the
Herzig et al., 1970; Zamani and Maini, 2009). Therefore, the PB combined effects of FL , FD , FB , and FAB , which control the motion state of
mechanism during tailwater reinjection in the WCSGR can be described colloidal particles. FE can be used to analyze the adsorption and
as follows: large particles exceeding 10 μm in size are prone to precip­ desorption process of 3 μm and 5 μm fluorescent microspheres. More­
itation or mechanical interception on the surface of sand grains due to over, Fw , Fs , FEy , and G are utilized to assess whether particles undergo
size filtering; Intermediate-sized particles ranging from 3 μm to 10 μm separation, while Fd and Fp are the main driving forces behind particle
are transported, but may also form bridge blockages or deposit inside motion.
sand grains with the combined action of physicochemical and hydro­ The movable sand grains and the injected fluorescent microspheres
dynamic effects; While colloidal particles smaller than 3 μm primarily in the pores of sandstone experience electrostatic force, hydrodynamic
exhibit absorption on the surface of sand grains or aggregation due to force, lift force, and gravity, in which hydrodynamic force and lift force
physicochemical effects, i.e., van der Waals forces, electric double layer work to separate particles from each other, while the electrostatic force
forces, Born repulsion, ect. A schematic diagram illustrating the and gravity make particles adhesion. Particle migration occurs when the
[( ) ]
blockage mechanism is shown in Fig. 15. Nevertheless, particle move­ adhesion torque FEy + G i between particles is smaller than the sep­
ment in a porous medium is complex, unpredictable, and difficult to [ ( ) ]
aration torque (Fw + Fs )i + Fd + Fp + FEx r . The fluid velocity, solu­
control, which greatly increases the difficulty of its research, but the tion salinity, pH, suspended particle size, and reservoir temperature may
results of this paper can provide valuable insights for further in-depth affect the particle transport state. An increase in fluid velocity during
study of particle motion. reinjection leads to higher flow drag and lift forces, potentially
enhancing the separation torque and promoting particle movement.
4.2. Physical model of particle motion Decreasing fluid salinity reduces the electrostatic force, resulting in a
decrease in adhesion torque and particles movement. Increases in tem­
Particle seepage with water flow in a sandstone medium involves the perature and pH can also weaken the electrostatic force, and the parti­
complex transfer of mass, momentum, and energy within a intricate cles may also move. Additionally, the direction of fluid action on
geometric structure. This process encompasses interactions between particles and the complex pore throat structure of sandstone GR intro­
particles themselves, particle-sand surface interactions, and particle- duce uncertainty in particle movement patterns, leading to an extremely
fluid interactions, primarily through migration and adsorption. The

Fig. 15. Schematic diagram of particle blockage.

10
J. Xia et al. Journal of Hydrology 624 (2023) 129981

Fig. 16. Force analysis of microspheres.

complex process of particle dispersion and release. As shown in Fig. 17, 4.3. Particle blockage evolution
the injected fluorescent microspheres can pass successfully through the
sandstone pore throats in the initial stage, but their transport to an in­ The filtration accuracy of tailwater in a recharge project can reach
ternal position is retained because of their large sizes. Alternatively, below 5 μm, while the pore throat sizes of a WCSGR is relatively large.
movable sand within the sandstone may directly block pore throats Therefore, the development process of PB during tailwater recharge can
under fluid movement, resulting in permanent blockage at specific lo­ be roughly divided into three periods in terms of the influence mecha­
cations. In conclusion, the mechanism of PB within sandstone pores nism of microparticles on the core permeability: the initial, middle and
mainly includes these effects, i.e., size filtering, hydraulics, inertia, later periods. Adsorption dominates in the initial period of recharge
diffusion, and sedimentation. seepage. In this period, the adsorption of particles is strong, while the
washing and drag forces of the fluid on particles are not sufficient.
Therefore, particles carried by the water flow easily deviate from the
flow line and constantly settle on the surface of the rock samples to be

Fig. 17. Motion model of microspheres.

11
J. Xia et al. Journal of Hydrology 624 (2023) 129981

adsorbed. This accumulation causes a rapaid decline in surface porosity Bureau, Henan Province and Sinopec Xinxing New Energy Research
and permeability of the samples. In the middle period, desorption be­ Institute Co., Ltd. for providing rock samples for this study.
comes the main effect, and the particles adsorbed and deposited on the
surface of the samples increase continuously, which makes the available References
cross-sectional area for water flow decrease gradually. Consequently,
the pore flow velocity increase, along with the hydrodynamic shear Ahfir, N.-D., Benamar, A., Alem, A., Wang, HuaQing, 2009. Influence of Internal
Structure and Medium Length on Transport and Deposition of Suspended Particles: A
stress. Then, the washing and drag forces on particles are enhanced,
Laboratory Study. Transport in Porous Media 76 (2), 289–307.
causing some particles to detach from the sample surface and then Ahfir, N.-D., Hammadi, A., Alem, A., Wang, HuaQing, Le Bras, G., Ouahbi, T., 2017.
migrate to the interior of the core under the action of hydrodynamic Porous media grain size distribution and hydrodynamic forces effects on transport
and deposition of suspended particles. Journal of environmental sciences 53,
forces. In the later period, particle migration and blockage occur under
161–172.
the joint action of pore throat characteristics, particle size, and hydro­ Anderson, A., Rezaie, B., 2019. Geothermal technology: Trends and potential role in a
dynamic forces of the sample until the permeability of the entire sample sustainable future. Applied Energy 248, 18–34. https://doi.org/10.1016/j.
decreases to a stable value. apenergy.2019.04.102.
Badalyan A. et al., 2015. Laboratory Study on Formation Damage in Geothermal
Reservoirs Due to Fines Migration.
5. Conclusions and prospects Bai B., Xu T., Guo Z., 2016. An experimental and theoretical study of the seepage
migration of suspended particles with different sizes. Hydrogeology Journal. 24.
2063-2078.
In this study, experimental phenomena, theoretical analysis, and Bai, B., Long, F., Rao, D., Xu, T., 2017. The effect of temperature on the seepage transport
empirical criteria were used to explore the mechanism and mode of of suspended particles in a porous medium. Hydrological Processes 31 (2), 382–393.
particle blockage (PB) in a weakly consolidated sandstone geothermal https://doi.org/10.1002/hyp.11034.
Bennacer, L., Ahfir, N.-D., Alem, A., Wang, HuaQing, 2017. Coupled Effects of Ionic
reservoir (WCSGR). Strength, Particle Size, and Flow Velocity on Transport and Deposition of Suspended
The core permeability declines exponentially over time, following a Particles in Saturated Porous Media. Transport in Porous Media 118 (2), 251–269.
double exponential decay function. Microspheres injected at a turbidity Bizmark, N., Schneider, J., Priestley, R.D., Datta, S.S., 2020. Multiscale dynamics of
colloidal deposition and erosion in porous media. Sci Adv 6 (46).
of 30 mg/L exhibit higher accumulation on the sample surface compared
Bouwer, H., 2002. Artificial recharge of groundwater: hydrogeology and engineering.
to 20 mg/L and 10 mg/L. However, the flow rate of 4 mL/min at the Hydrogeol J 10 (1), 121–142.
same turbidity can slow the increase in the density of its aggregation on Caselles-Osorio, A., Puigagut, J., Segú, E., Vaello, N., Granés, F., García, D., García, J.,
2007. Solids accumulation in six full-scale subsurface flow constructed wetlands.
the surface. The experimental condition of 25 ◦ C causes the least amount
Water Research 41 (6), 1388–1398.
of damage to the core permeability compared to other temperature Chai R. et al., 2022. Formation damage of sandstone geothermal reservoirs: During
conditions. decreased salinity water injection. Applied Energy. 322. 119465. DOI:https://doi.
Particles larger than 10 μm precipitated or are mechanically inter­ org/10.1016/j.apenergy.2022.119465.
Chen X., Bai B., 2014. Experimental investigation and modeling of particulate
cepted on the surface of sand grains by size filtering, and intermediate transportation and deposition in vertical and horizontal flows. Hydrogeology
particles between 3 μm and 10 μm are transported, forming bridge Journal. 23. 365-375.
blocking or deposition within sand grains with the combined action of Chequer, L., Vaz, A., Bedrikovetsky, P., 2018. Injectivity decline during low-salinity
waterflooding due to fines migration. Journal of Petroleum Science and Engineering
physicochemical and hydrodynamic effects, while colloidal particles 165, 1054–1072. https://doi.org/10.1016/j.petrol.2018.01.012.
smaller than 3 μm mainly adsorbed on the surface of sand grains or Cui, X., Fan, Y., Wang, H., Huang, S., 2019. Experimental investigation of suspended
aggregated due to physicochemical effects. Then, the process of PB in particles transport in porous medium under variable temperatures. Hydrological
Processes 33 (7), 1117–1126.
the actual recharge project was described three stages: adsorption Cui, G., Ning, F., Dou, B., Li, T., Zhou, Q., 2022. Particle migration and formation
dominates in the initial stage, desorption is prominent in the middle damage during geothermal exploitation from weakly consolidated sandstone
stage, and particle migration/blockage occurs in the later stage. reservoirs via water and CO2 recycling. Energy 240, 122507. https://doi.org/
10.1016/j.energy.2021.122507.
Therefore, in pursuing the high filtering accuracy of the tailwater, Dillon, P., Pavelic, P., Massmann, G., Barry, K., Correll, R., 2001a. Enhancement of the
the geothermal recharge technology of the WCSGR should consider membrane filtration index (MFI) method for determining the clogging potential of
increasing the ion concentration of the tailwater, reducing its particle turbid urban stormwater and reclaimed water used for aquifer storage and recovery.
Desalination 140 (2), 153–165. https://doi.org/10.1016/S0011-9164(01)00365-4.
turbidity, optimizing the flow rate, and maintaining the temperature of
Dillon, P., Pavelic, P., Massmann, G., Barry, K., Correll, R., 2001b. Enhancement of the
the tailwater recharge at approximately room temperature for economic membrane filtration index (MFI) method for determining the clogging potential of
dfficiency. turbid urban stormwater and reclaimed water used for aquifer storage and recovery.
Nevertheless, this paper has the following deficiencies that can be Desalination 140, 153–165. https://doi.org/10.1016/S0011-9164(01)00365-4.
Frey, J.M., Schmitz, P., Dufreche, J., Gohr Pinheiro, I., 1999. Particle Deposition in
subsequent research directions: incorporating numerical simulations, Porous Media: Analysis of Hydrodynamic and Weak Inertial Effects. Transport in
considering the inpact of particle shapes, and optimizing instrumenta­ Porous Media 37 (1), 25–54. https://doi.org/10.1023/A:1006546717409.
tion for real-time dynamic monitoring of the flow process. Glover, P.W., Walker, E., 2009. Grain-size to effective pore-size transformation derived
from electrokinetic theory. GEOPHYSICS 74 (1), E17–E29. https://doi.org/10.1190/
1.3033217.
Declaration of Competing Interest Goldman, A.J., Cox, R.G., Brenner, H., 1967. Slow viscous motion of a sphere parallel to
a plane wall—II Couette flow. Chemical Engineering Science 22 (4), 653–660.
https://doi.org/10.1016/0009-2509(67)80048-4.
The authors declare that they have no known competing financial Goss, D.W., Smith, S.J., Stewart, B.A., Jones, O.R., 1973. Fate of suspended sediment
interests or personal relationships that could have appeared to influence during basin recharge. Water Resour Res 9 (3), 668–675. https://doi.org/10.1029/
the work reported in this paper. WR009i003p00668.
Han, X., Zhong, L., Liu, Y., Fang, T., Chen, C., 2020. Experimental Study and Pore
Network Modeling of Formation Damage Induced by Fines Migration in
Data availability Unconsolidated Sandstone Reservoirs. Journal of Energy Resources Technology-
transactions of The Asme 142.
Heiland, J.C., Flor, M.E., 2006. Influence of Rock Failure Characteristics on Sanding
Data will be made available on request.
Behavior: Analysis of Reservoir Sandstones from the Norwegian Sea. SPE
International Symposium and Exhibition on Formation Damage Control. https://doi.
Acknowledgments org/10.2118/98315-ms.
Herzig, J.P., Leclerc, D.M., Goff, P.L., 1970. Flow of Suspensions through Porous
Media—Application to Deep Filtration. Industrial & Engineering Chemistry 62 (5),
Funding: This research was funded by The National Key Research 8–35. https://doi.org/10.1021/ie50725a003.
and Development Programs of China 2019YFB1504204, Hu, W., 2012. Clogging Mechanism and Prevention Measures Research of Geothermal
2019YFB1504201, and 2019YFB1504203. Water Reinjection-Take Xianyang Urban Area for example. Chang’an University.
Master Thesis.
The authors thank the Second Geological and Mineral Exploration Hu, Q., Brusseau, M.L., 1994. The effect of solute size on diffusive-dispersive transport in
Institute of Geological and Mineral Exploration and Development porous media. Journal of Hydrology 158 (3-4), 305–317.

12
J. Xia et al. Journal of Hydrology 624 (2023) 129981

Kamila, Z., Kaya, E., Zarrouk, S.J., 2021. Reinjection in geothermal fields: An updated sandstone. Geothermics 53, 225–235. https://doi.org/10.1016/j.
worldwide review 2020. Geothermics 89, 101970. https://doi.org/10.1016/j. geothermics.2014.06.004.
geothermics.2020.101970. Song, W., Liu, X., Zheng, T., Yang, J., 2020. A review of recharge and clogging in
Khilar K.C., Fogler H.S., 1998. Migrations of Fines in Porous Media. Migrations of Fines sandstone aquifer. Geothermics 87, 101857.
in Porous Media. Song, W., Liu, X., Zheng, C., Wang, H., 2021. Migration-deposition Characteristics of
Khilar, K.C., Fogler, H.S., 1984. The existence of a critical salt concentration for particle Exogenous Particles Near the Injection Well in a Groundwater Heat Pump System.
release. J Colloid Interf Sci 101 (1), 214–224. https://doi.org/10.1016/0021-9797 Geothermics 94, 102097.
(84)90021-3. Torkzaban, S., Bradford, S.A., Vanderzalm, J.L., Patterson, B.M., Harris, B., Prommer, H.,
Kia, S.F., Fogler, H.S., Reed, M.G., Vaidya, R.N., 1987. Effect of Salt Composition on Clay 2015. Colloid release and clogging in porous media: Effects of solution ionic strength
Release in Berea Sandstones. SPE Production Engineering 2 (04), 277–283. https:// and flow velocity. Journal of Contaminant Hydrology 181, 161–171.
doi.org/10.2118/15318-pa. Tripathy, A., 2010. Hydrodynamically and chemically induced in situ kaolin particle
Kim, S.-J., Benefield, L.D., 1996. Physicochemical Factors Influencing Colloidal Particle release from porous media an experimental study. Advanced Powder Technology 21
Transport in Porous Media. Separation Science and Technology 31 (19), 2621–2653. (5), 564–572. https://doi.org/10.1016/j.apt.2010.02.012.
https://doi.org/10.1080/01496399608000817. Vigneswaran, S., Suazo, R.B., 1987. A detailed investigation of physical and biological
Kong, Y., Pang, Z., Shao, H., Hu, S., Kolditz, O., 2014. Recent studies on hydrothermal clogging during artificial recharge. Water, Air, and Soil Pollution 35 (1), 119–140.
systems in China: a review. Geothermal Energy 2 (1), 19. https://doi.org/10.1186/ https://doi.org/10.1007/BF00183848.
s40517-014-0019-8. Wang, Y., Yu, M., Bo, Z., Bedrikovetsky, P., Le-Hussain, F., 2021a. Effect of temperature
Liu, Q., Cui, X., Zhang, C., Huang, S., 2016. Experimental investigation of suspended on mineral reactions and fines migration during low-salinity water injection into
particles transport through porous media: particle and grain size effect. Berea sandstone. Journal of Petroleum Science and Engineering 202, 108482.
Environmental Technology 37 (7), 854–864. https://doi.org/10.1016/j.petrol.2021.108482.
Liu, X., Wang, Y., Li, S., Jiang, X., Fu, W., 2020. The influence of reinjection and Wang, Y., Yu, M., Bo, Z., Bedrikovetsky, P., Le-Hussain, F., 2021b. Effect of temperature
hydrogeological parameters on thermal energy storage in brine aquifer. Applied on mineral reactions and fines migration during low-salinity water injection into
Energy 278, 115685. https://doi.org/10.1016/j.apenergy.2020.115685. Berea sandstone. Journal of Petroleum Science and Engineering 202, 108482.
Lun Z., 2006. Study on the blockage characteristics of reservoir pores due to suspension Yan X., et al., 2019. Sand Prediction of Medium Strength Sandstones in Offshore
particles of injection water. Journal of Logistical Engineering University (03). 30-32 Oilfields. Journal of Yangtze University(Natural Science Edition). 16 (04). 44-48+
+ 39. 46.
McCarthy, J., Zachara, J., 1989. ES&T Features: Subsurface transport of contaminants. Yang, Y., Yuan, W., Hou, J., You, Z., 2022. Review on physical and chemical factors
Environmental Science & Technology 23 (5), 496–502. https://doi.org/10.1021/ affecting fines migration in porous media. Water Research 214, 118172. https://doi.
es00063a602. org/10.1016/j.watres.2022.118172.
Muneer, R., Hashmet, M.R., Pourafshary, P., 2020. Fine Migration Control in Sandstones: You, Z., Badalyan, A., Yang, Y., Bedrikovetsky, P., Hand, M., 2019. Fines migration in
Surface Force Analysis and Application of DLVO Theory. ACS Omega 5 (49), geothermal reservoirs: Laboratory and mathematical modelling. Geothermics 77,
31624–31639. https://doi.org/10.1021/acsomega.0c03943. 344–367. https://doi.org/10.1016/j.geothermics.2018.10.006.
Ochi, J., Vernoux, J.-F., 1998a. Permeability decrease in sandstone reservoirs by fluid Yu, L., Zhang, L., Zhang, R., Ren, S., 2018. Assessment of natural gas production from
injection: Hydrodynamic and chemical effects. Journal of Hydrology 208 (3), hydrate-bearing sediments with unconsolidated argillaceous siltstones via a
237–248. https://doi.org/10.1016/S0022-1694(98)00169-3. controlled sandout method. Energy 160, 654–667. https://doi.org/10.1016/j.
Ochi, J., Vernoux, J.F., 1998b. Permeability decrease in sandstone reservoirs by fluid energy.2018.07.050.
injection Hydrodynamic and chemical effects. J Hydrol 208 (3–4), 237–248. Zamani, A., Maini, B., 2009. Flow of dispersed particles through porous media — Deep
Oliveira, M.A., Vaz, A.S.L., Siqueira, F.D., Yang, Y., You, Z., Bedrikovetsky, P., 2014. bed filtration. Journal of Petroleum Science and Engineering 69 (1), 71–88. https://
Slow migration of mobilised fines during flow in reservoir rocks: Laboratory study. doi.org/10.1016/j.petrol.2009.06.016.
Journal of Petroleum Science and Engineering 122, 534–541. Zhang L. et al., 2020. Particle migration and blockage in geothermal reservoirs during
O’Neill, M.E., 1968. A sphere in contact with a plane wall in a slow linear shear flow. water reinjection: Laboratory experiment and reaction kinetic model. Energy. 206:
Chemical Engineering Science 23 (11), 1293–1298. https://doi.org/10.1016/0009- 118234. DOI:https://doi.org/10.1016/j.energy.2020.118234.
2509(68)89039-6. Zhang L. et al., 2022b. Formation blockage risk analysis of geothermal water reinjection:
Platzer, C., Mauch, K., 1997. Soil clogging in vertical flow reed beds - Mechanisms, Rock property analysis, pumping and reinjection tests, and long-term reinjection
parameters, consequences and solutions? Water Sci Technol 35 (5), 175–181. prediction. Geoscience Frontiers. 13 (1). 101299. DOI:https://doi.org/10.1016/j.
Rehg, K.J., Packman, A.I., Ren, J., 2005. Effects of suspended sediment characteristics gsf.2021.101299.
and bed sediment transport on streambed clogging. Hydrological Processes 19 (2), Zhang, L., Geng, S., Chao, J., Wen, R., Yang, L., Ren, S., 2022a. Evaluation of Formation
413–427. https://doi.org/10.1002/hyp.5540. Blockage Risk in Geothermal Water Reinjection Using a Novel Particle Migration
Rosenbrand, G., Fabricius, I.L., 2014. Effect of Hot Water Injection on Sandstone Model Based on the Bridging Principle. Arabian Journal for Science and Engineering
Permeability - An Analysis of Experimental Literature (SPE 154489). Geothermics 50 47 (9), 11807–11823.
(2), 155–166. Zhao, J., Ye, X., Du, X., Zhang, X., Cui, R., 2019. Study on the migration characteristics of
Rosenbrand, E., Kjøller, C., Riis, J.F., Kets, F., Fabricius, I.L., 2015. Different effects of suspended particles inporous media during different concentration of suspension.
temperature and salinity on permeability reduction by fines migration in Berea Water Resources and Hydropower Engineering 50 (10), 25–31.

13

You might also like