You are on page 1of 24

Journal of Petroleum Science and Engineering 207 (2021) 109147

Contents lists available at ScienceDirect

Journal of Petroleum Science and Engineering


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

Investigation of sand production prediction shortcomings in terms of


numerical uncertainties and experimental simplifications
Mohammad Hossein Shahsavari a, Ehsan Khamehchi a, *, Vahidoddin Fattahpour b,
Hamed Molladavoodi c
a
Amirkabir University of Technology (Tehran Polytechnic), Department of Petroleum Engineering, Tehran, Iran
b
University of Tehran, School of Mining Engineering, Tehran, Iran
c
Amirkabir University of Technology (Tehran Polytechnic), Department of Mining Engineering, Tehran, Iran

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

Keywords: Sand production is one of the main research topics in the petroleum industry. This problematic phenomenon is
Sand production related to the mechanical and hydrodynamic behavior and reservoir specifications. Sand production often leads
Numerical simulation to equipment damage and significant losses. This is usually studied by experimental and numerical methods.
Experimental
This study first gives a brief overview of the methods used to predict sand production in terms of experimental
Uncertainties
tests, numerical simulation, and field data. The main objective of this research is to highlight the shortcomings of
Errors
these methods. In addition, experimental hollow cylinder test data and widely used continues numerical simu­
lation are used to investigate these shortcomings. This study is performed by scrutinizing the most important and
basic parameters in numerical modeling including two constitutive models (Mohr-Coulomb elastic-perfectly
plastic and Mohr-Coulomb cohesion softening/friction hardening failure criteria) and various element sizes
and shapes. Since most studies use continuum approaches, it was decided to use a finite difference program.
Further, to reduce the undesirable influencing factors the fracture energy regularization method was imple­
mented to diminish mesh dependency related to energy dissipation. In addition, a mesh size sensitivity analysis
was performed to show the effects of size, shape and pattern of mesh on the results and cumulative probability
distribution versus absolute relative error diagrams were applied to compare the accuracy of the models. After
all, the best model of prediction was selected to simulate a real sand production in an oil well with 50◦ incli­
nation and 6 SPF in North Sea Reservoir.
Review of experimental papers shows that these tests usually have several hypotheses for simplification. Rock
strength, stress state, fluid flow properties, test duration, and sample dimensions are the most important pa­
rameters in sand production tests. Researchers usually focus on only one or two of these parameters, which may
lead to many errors in contrast to reality. In addition, numerical methods have some deficiencies such as mesh
size problems, lack of critical-state-based constitutive models, sanding criteria, and lack of sufficient research for
the calibration procedure in the DEM model. A small change in both the numerical simulation input data and the
experimental procedure leads to different results. Indeed, the experimental test of sand production with multiple
hypotheses can qualitatively simulate the steps and shape of well or perforation failure not the exact sand
production rate. Also, numerical simulation results are very sensitive due to uncertainties and errors of the model
and input data. It has been shown that the results are highly dependent on every simple parameter such as the
shape of the elements. The differences between experimental results and numerical modeling in only one
perforation can be 2 to 10 times. Despite all errors and uncertainties in both the laboratory and modeling studies,
the simulation of a real sand production in North Sea Reservoir with same sized element was relatively
acceptable. This level of accuracy of the model can be very helpful in subsequent decision making and sand
control management.

* Corresponding author.
E-mail address: khamehchi@aut.ac.ir (E. Khamehchi).

https://doi.org/10.1016/j.petrol.2021.109147
Received 22 February 2021; Received in revised form 12 May 2021; Accepted 20 June 2021
Available online 27 June 2021
0920-4105/© 2021 Elsevier B.V. All rights reserved.
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

1. Introduction include: 1. Extremely complex nature of sand production, 2. Weaknesses


and shortcomings of analytical, numerical, and experimental methods,
Nearly two-thirds of the world’s petroleum reservoirs are found in 3. Multiplicity of field data and measurement uncertainty 4. Lack of a
unconsolidated sandstone (Osisanya, 2010) and the probability of sand comprehensive benchmark for sand production criteria (Morita, 1994).
production in these formations is very high. This phenomenon affects The complex nature of sand production is originated from the presence
hydrocarbon production rates, field economics, continuous production, of many various parameters that are very difficult to study them
as well as safety. This problem has caused researchers to seek solutions together in experimental setups under different conditions of wellbore.
over the decades to mitigate sand production in the petroleum industry Also, when modeling sand production, accuracy, complexity, run time,
(Aborisade, 2010). Moreover, produced solids from hydrocarbon res­ necessity of several input data, the shape, size, and number of elements,
ervoirs are hazardous to the environment. Millions of dollars are spent engineering judgment of the modeler all affect the results.
every year to prevent sand production problems and repair adverse ef­ Absence of a comprehensive benchmark for sand production criteria
fects (Ikporo and Sylvester, 2015). Consequently, there is a need to is due to different failure patterns and cavity shapes (spiral shear band,
address issues related to sand production prediction and prevention. V-shape breakout, dog-ear breakout, and slit mode cavity), hydrody­
Generally, sand production is investigated by four basic approaches namic and chemical effects, erosion, etc. In different rocks and stress
including analytical methods, empirical approaches using field obser­ states (Wang and Sharma, 2018). In addition, most numerical and
vations and offset well data, experimental tests, numerical approaches, analytical methods consider a two-dimensional model, which includes
and two or more methods are combined for more reliable prediction many simplifications and assumptions. On the other hand, the 3D
(Qiu et al., 2006; Subbiah, 2020). models have several shortcomings, such as long runtime, considering
Due to their simplicity, analytical methods have gained popularity in one cylinder perforation, and neglecting interactions among them due to
the oil and gas industry based on modeling of perforation cavity sta­ the limitation of the number of meshes (Rahmati et al., 2013). Discrete
bility. In these methods, producer well or perforation stability is related element model (DEM) methods can also be useful tools to elucidate the
to the stresses applied to the near-wellbore formation under producing mechanism of sand production, but they have several defects. In most
conditions. Researchers such as Risnes et al. (1982) investigated the DEM models, the micro properties cannot be evaluated by direct mea­
stresses around the wellbore using elastic-plastic Mohr-Coulomb theory. surement and it is necessary to calibrate them by macro responses via
Geertsma (1985) studied some geomechanical aspects of well comple­ slow and difficult process of laboratory tests. In addition, the size of the
tions and sand production in wells during injection and production. model that can be practically computed is limited due to the computa­
Simplifications are the most critical flaw of this method which does not tional cost (Climent Pera, 2016).
take into account all effective parameters. According to published papers, the predictive precision of available
The purpose of empirical methods is to establish a relationship be­ models is in the range of 0.1–10 times the real rate of sand production
tween one parameter or number of parameters including porosity, (Fuh and Nozaki, 2014; Kim and Sharma, 2012; van den Hoek and
drawdown, and rock stresses (Aborisade, 2010). Stein and Hilchie Geilikman, 2005; Wu et al., 2016). These numerical models suffer from
(1972) developed a method of the sand production rate prediction based measurement and model errors. The uncertainties in the input data and
on evaluation formation strength from log data and offset well condi­ models can greatly affect the reliability of sand production numerical
tion. Khamehchi and Reisi (2015) used shear modulus to bulk model (Veeken et al., 1991).
compressibility relation for sand production prediction with field ob­ Properties such as permeability, porosity, viscosity, rock strength,
servations and well data. Also, some researchers such as Abbas et al. Poisson’s ratio and in situ stresses, etc., are uncertainties (Acock et al.,
(2019) conducted sand production prediction analyses by developing a 2004). Some uncertainty variables can be changed during oil and gas
method that uses log data, measurable experimental data, analytical production. Other parameters such as wellbore radius and completion
relationships, and empirical methods together. One of the defects of this type are the result of design choices. Some parameters, such as skin, are
approach is the field-dependent results, which are different for each also an uncertainty. Nevertheless, the range of skin varies based on
field. completion type (Wehunt, 2003). If these uncertainties of the model are
Many authors have studied different numerical methods under ignored in numerical simulation, the model predictions and thus the
changing down-hole conditions (Morita et al., 1989). In the past few decisions based on the prediction results could be biased.
decades, several sand prediction models with different levels of Furthermore, in sand production experimental test there are many
complexity have been developed based on various numerical simulation uncertain parameters and different conditions include effect of inner and
methods. Numerical models incorporate the full range of rock behavior outer hole diameter, different combinations of fluid rate and stress
during elastic, plastic, and time-dependent deformation, as well as a value, radial and axial stress ratio, and real or synthetic sample. These
detailed description of the stress state. Numerical modeling is one of the uncertainties in the experimental parameters lead to results with low
most common prediction methods because it takes into account more reliability.
factors influencing rock failure and sand production (Bratli and Risnes, These errors are likely to be exacerbated by the combination of
1981; Coates and Denoo, 1981; Geilikman et al., 1994; Nouri et al., laboratory and modeling results. Table 1 shows sand production nu­
2007; Subbiah, 2020; Weingarten and Perkins, 1995). The merical model studies performed by some researchers. As can be seen,
time-consuming and costly implementation are the main drawback of the differences between modeling and experimental results in some
this method. models are very large. According to published papers, the prediction
The experimental method of sand production is the best way to un­ accuracy of the available models is in the range of 0.1–10 times the real
derstand the complex nature of this phenomenon. Generally, a hollow rate of sand production. Large differences in 3D models may be due to
cylinder core is used to simulate the sand production (Veeken et al., radial variations in the size of elements, while in 2D models using the
1991; Vriezen et al., 1975). The results of these experimental tests are a same size square elements this difference is smaller. It is worth noting
kind of graph (rate and cumulative produced sands) and a series of that these studies were conducted in a short period of time and no sig­
images of the hole after failure. The setup of these tests consists of the nificant research has been done in the long-term to study time depen­
triaxial cell, hollow cylinder samples, sand detection, and measurement dent parameters.
unit or a simple orifice to simulate the perforation sand arching. This Comparing the trends in the number of sand production prediction
method has some obstacles such as rock type selection, perforation size, and sand production prevention papers can manifest the weakness of
fluid characteristics, stress and pressure applying and etc. Which causes sand predictions methods in another way. In the last four decades, ac­
many simplifications and assumptions in the model. cording to the trend of the number of sand production prediction and
In sand production prediction studies there are some difficulties sand production prevention papers (Fig. 1), it can be seen that the

2
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

Table 1 in detail in the next section in the form of a review. Then, a hollow
Some sand production numerical model studies carried out by (Chin and Ramos, cylinder sand production experiment test and a simple widely used
2002; Kim and Sharma, 2012; Nouri et al., 2004; Papamichos et al., 2001; numerical simulation are presented to show sand production numerical
Rahmati et al., 2012). model’ deficiencies. This is done by scrutinizing the most important and
Reference Description Experiment to numerical basic parameters in numerical modeling including the constitutive
rate models (Mohr-Coulomb elastic-perfectly plastic and Mohr-Coulomb
Papamichos et al. Finite element-3D Poro- Tests in different external cohesion softening/friction hardening failure criteria) and different
(2001) elastoplastic stress element sizes and shapes. Since most studies use continuum approaches
Erosion Test 1: 1.25 we decided to use a finite difference program whose governing equa­
Test 2: 0.63
Test 3: 1.08
tions include the continuity, momentum, and Darcy equations. In
Chin and Ramos Finite element-3D strain- Tests in different pressure addition, a number of methods such as the fracture energy regulariza­
(2002) based model and flow rate tion method were used to diminish mesh dependency related to energy
single-phase, fully coupled Test 1: 0.87 dissipation.
Test 2: 14.58
Also, error analysis in terms of the cumulative probability distribu­
Test 3: 1.45
Test 4: 0.48 tion versus the absolute relative error helps to compare the prediction
(Nouri et al., 2004) Finite Difference-2D 0.68 performance of different methods. In addition, real sand production in
Erosion an oil well with an inclination of 50◦ was simulated in North Sea
Bilinear hardening/softening Reservoir to survey the accuracy of final results and to compare them
Mohr-Coulomb
with the field data.
Rahmati et al. Finite Difference-2D 1.38
(2012) Erosion
Strain hardening/softening 1.1. Experimental test
Mohr-Coulomb
Kim and Sharma Finite Difference-3D 1.08
(2012) Mohr-Coulomb two-phase
In sand production experimental test there are many uncertain pa­
rameters, different conditions, and various assumptions in laboratory
research, including reservoir scaling down, working with synthetic
samples or sieved river sandstone, and limited combinations of two-
phase flows and stress values, the radial and axial stress ratio and the
effect of inner and outer hole diameter. These uncertainties in experi­
mental parameters lead to results with low reliability. In addition, some
parameters such as pore pressure, temperature, and number of fluid
phases, rock mechanical parameters and testing time may affect the final
results. Ranjith et al. (2014) comprehensively studied the setup
arrangement of sand production cell improvement over time. The
simplest configuration was Terzaghi (1936)’s model, which includes a
box filled with sand with a trap door at the bottom. Indeed, the first
models examine sand arching at the perforation. The primary experi­
mental setups designed quantitatively with a simple orifice and single
fluid conduction. Over time, apparatuses were developed to perform
qualitative and quantitative tests, radial, axial and multiphase fluids in
uniaxial and tri-axial stress state.
Researchers usually examine one or two effective parameters in
laboratory tests. The most important effective parameters for sand
Fig. 1. Number of publications about sand production prediction and preven­ production process can be categorized into four main groups (Table 2)
tion in the period 1980–2020. (Ranjith et al., 2014).
Table 3 shows the most important experimental tests for sand pro­
number of prediction paper has an almost steady trend. Although, duction and their test conditions. The main purpose of these tests is
published papers in relation to prevention methods have an upward usually to identify the nature and stages of sand production and to find a
trend from 1980 to 2020. relationship between the production of sands and the influencing pa­
The installation of sand controllers in problematic wells is inevitable rameters such as pressure, fluid type, and rock material.
and therefore research into prevention is more appealing. Despite the Approximations are the main part of laboratory experiments and this
growth of new technologies and powerful computers, sand production fact is used in the analysis of sand production. With these assumptions,
modeling still has many problems due to the complexity of this phe­ the results are subject to doubt when applied to field-scale studies. For
nomenon. With this explanation, designing a sand control tool with the example, scaling down made in laboratory setups are usually affected by
specified shape and size is much easier than modeling the prediction of boundary effects (small sample problems and their stress concentrations
sand production due to the complexity and multiplicity of effective pa­ in comparison with the infinitely-applied stresses in the oil field).
rameters and variables. Generally, for managing sand production in oil
and gas field, a simplified prediction method along with well-known Table 2
prevention tools are used. The most important effective parameters for the sand production process.
Although abundant efforts has been put by many practitioners
Most important effective Examples
related to experimental and numerical modeling work, there are still parameters
some significant gaps in knowledge that need to be addressed (Subbiah,
Fluid flow Fluid flow rate, fluid phase, fluid compounds,
2020).
viscosity, flow direction and saturation
In this study, the complexity and shortcomings of sand production Rock material Grain size, shape, roughness, cementation,
prediction in terms of errors, limitations and uncertainties in field data, characteristics compactness and strength of the geological formation
experimental tests and numerical calculations are discussed descrip­ Loading effect Stress value and direction, and boundary effect
tively. These three methods of sand production prediction are discussed Outlet size Hole size, Particle size

3
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

Table 3
The most important experimental tests for sand production and their test conditions.
Model Geometry Phase Rock P (psi) E (GPa) Failure Failure Time Goals
pattern reasons of test
T (C) Poison
(min)

Zivar et al. (2019) Cylinder synthetic Synthetic 1400 N/A complete mechanical 12 Investigate sand
5.3 cm (H) reservoir unconsolidated 21C N/A collapse and failure production stages
3.8 cm (D) brine sandstone failure of and permeability
the core alterations during
sand production
Deng et al. (2019) Hollow gas Unconsolidated 30 1.5 N/A mechanical 230 Study the critical
Cylinder Sandstone 21C 0.25 failure production pressure
6.6 cm (H) Reservoir difference and the
3.7 cm (D) critical flow rate of
1.8 cm (perf the sand production
D) at unconsolidated
3.3 cm (perf sandstone reservoir
H) in a deep-water gas
field
Lv (2018) Hollow water Gosford 10,000 13.73 N/A mechanical N/A The experiments
Cylinder sandstone 21C 0.18 failure done under different
6–8 cm (H) compression
3–4.1 cm scenarios for
(D) validation purposes
0.5 cm (perf and Digital Image
D) Correlation and 3D
1 cm (perf X-Ray Computed
H) Tomography.
Wu et al. (2016) Hollow Single Synthetic and 7000 0.55–10 Shear bands mechanical 2–3 Real-time
cylinder phase natural 21C 0.1–0.34 failure week monitoring using X-
31 cm (H) Oil sandstone tensile ray CT scanning
20 cm (D) Gas failure
1.8 cm (perf
D)
Meier et al. (2013) Hollow NA Posidonia shale 1400–40,000 5 breakout erosion NA Influence of
cylinder 21C N/A mechanical borehole diameter
2–8 cm (H) failure on the formation of
1–5 cm (D) borehole breakouts
0.1–1.9 cm
(perf D)
Ranjith et al. (2013) Cubic Air Unconsolidated 100 N/A Sand arches Drag force 60 Examine factors
2.5 × 25.5 Moisture sand (compacted 21C N/A collapse influencing sand
× 25.5 cm content sand) production from a
1-3-5 cm screen
(perfD)
Wu and Choi (2012) Hollow Brine MF sandstone 10,000 20 Break-outs erosion N/A Study fundamental
cylinder Oil CG sandstone 21C N/A Shear band mechanical mechanisms of sand
14 cm (H) slot type failure production-
8.6–20 cm cavity perforation failure
(D) development
0.16–1.9 cm The failure
(perf D) condition of a
10 cm (perf hollow cylinder was
H) investigated using a
different types of
geomechanical tests
Fattahpour et al. Hollow Oil Synthetic 500–2500 1.6–3.02 curved slabs Pure 24 An innovative
(2012) Cylinder Single sandstone 14.7 of failed mechanical method was used to
30 cm (H) phase @hole sand failure is the continuously
15 cm (D) Axial- 21C 0.18–0.19 breakouts reason for measure the
1-2z cm Radial sand produced sand.
(perf D) production Study the sand
15 cm (perf production in terms
L) of rock strength and
perforation size
Rahmati et al. (2012) Hollow Single Unconsolidated 0–7000 0.000575 N/A mechanical 1200 The test was
Cylinder phase SWS sandstone 21C 0.1–0.25 failure intended to simulate
25.4 cm (H) N/A and erosion sand production
15.24 cm from a perforation
(D) cavity
1.27 cm
(perf D)
5.08 cm
(perf L)
Choi (2011) Hollow Two unconsolidated 35,000 N/A N/A seepage 60 Perforating
Cylinder phase sandstone 21C N/A failure unconsolidated
28–45.5 cm sandstone and
(continued on next page)

4
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

Table 3 (continued )
Model Geometry Phase Rock P (psi) E (GPa) Failure Failure Time Goals
pattern reasons of test
T (C) Poison
(min)

(H) Brine and mechanical measuring the


10.7–17.78 kerosene failure amount of produced
cm (D) sand during
0.6–0.85- perforating and
1.25 cm post-shot flow
(perf D)
Choi (2011) Hollow Single Fine sand 14.7 0.4 N/A seepage 15 Investigation of
Cylinder phase 21C 0.43 failure mechanisms of sand
16.6 cm (H) odorless production and
107 cm (D) mineral mechanical
2.22 cm spirits characterization of
(perf D) (OMS) caverns created in
Axial- carbonate rock
Radial formations for
N/A natural gas storage
Nouri et al. (2006) Hollow Single cylinder 1800 0.4 N/A mechanical N/A Obtaining real time
Cylinder phase synthetic 21C 0.43 failure sand production
25 cm (H) odorless sandstone data
12.5 cm (D) mineral
2.54 cm spirits
(perf D) (OMS)
Axial-
Radial
Wu et al. (2005) Hollow Two Outcrops and 10,000 N/A Break-outs seepage N/A Investigate series of
cylinder phase cores sandstone 21C N/A Shear band failure experimental
14 cm (H) Water slot type mechanical perforation collapse
8.6–20 cm Oil-gas cavity failure tests in order to
(D) explain and quantify
0.16–1.9 cm the water-cut effect
(perf D) on cavity failure
10 cm (perf
H)
Haimson and Lee Cubic Drilling Tablerock 10,000 1.5–3.5 V-shaped mechanical 33 Compare the shape
(2004) 15 × 15 × fluid sandstone 21C 0.23–0.27 (dog-eared) failure and failure
23 cm Break-outs mechanism around
2.2 cm (perf vertical perforation
D) in high-porosity
sandstones
Chin and Ramos Hollow Single Salt Wash 7000 0.00001–0.00007 telltale seepage 400 The effect of rock
(2002) Cylinder phase Oil South (SWS) 21C 0.11 onion-like failure strength, flow rate,
60.96 cm Sandstone – layering mechanical oil viscosity, and
(H) 0.35 failure producing time of
38.1 cm (D) sand production
1.27 cm investigated.
(perf D)
33.02 cm
(perf L)
Papamichos et al. Hollow Single weak, 6000 N/A localized erosion 1895 Experimental
(2001) cylinder phase oil compacted, 21C N/A shear band mechanical volumetric sand
20 cm (H) synthetic, 21C N/A failure production data
10–20 cm sandstones measuring
(D)
2 cm (perf
D)
Walton et al. (2001) Hollow Two unconsolidated 35,000 N/A N/A seepage 60 Perforating
Cylinder phase sandstone 21C N/A failure unconsolidated
28–45.5 cm Brine and mechanical sands: an
(H) kerosene failure experimental
10.7–17.78 investigation.
cm (D) Measuring the
0.6–0.85- amount of produced
1.25 cm sand during
(perf D) perforating, as well
as that produced
during fluid flow8
Van den Hoek et al. Cubic Brine Artificial 2500 0.4 V-shaped mechanical 350 Long-term stability
(2000) 26.25 × OMS sandstone 21C 0.05 (dog-eared) failure of horizontal
26.25 × 38 Break-outs wellbore
25.4 cm completions.
(perf D) The influence of
rock failure in the
near-wellbore
region on well
(continued on next page)

5
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

Table 3 (continued )
Model Geometry Phase Rock P (psi) E (GPa) Failure Failure Time Goals
pattern reasons of test
T (C) Poison
(min)

productivity was
studied.
Al-Awad et al. (1999) Cylinder Brine unconsolidated 1200 N/A N/A mechanical N/A The experiments
8 cm (H) Light sandstone 21C N/A failure were designed to st
3.81 cm (D) crude oil formation the effect of
confining pressure,
flow rate, and fluid
viscosity on sand
production in
unconsolidated
sandstone
Ewy et al. (1999) Hollow Single Sandstone 6000 N/A V-shaped mechanical N/A Wellbore failure
Cylinder phase 21C 0.2 (dog-eared) failure was simulated in the
6.63 cm (H) Water Break-outs laboratory by
3.81 cm (D) subjecting a
1.27 cm perforation to
(perf D) applied in-situ
stresses
Unander et al. (1997) Hollow Single Weak Red 4000 2–3 Break-outs mechanical N/A Flow geometry
Cylinder phase Wildmoor 21C N/A failure effects on sand
15 cm (H) Oil sandstone (RW) and erosion production
10 cm (D)
2 cm (perf
D)
6 cm (perf L)
Behrmann et al. Cylinder Brine Castlegate 3500 N/A N/A mechanical 5000 Evaluation of sand
(1997) 69 × 69 × OMS sandstone 21C N/A failure production versus
81 cm confining stresses,
12 cm (hole flow rate and fluid
D) type.
5.58 cm
(perf D)
Tronvoll et al. (1997) Hollow Single Homogeneous 3000 N/A Surface seepage 120 Experimental study
Cylinder phase synthetic 21C N/A spalling failure of sand production
15 cm (H) sandstone Buckling, mechanical and the effect of oil
10 cm (D) shear band failure production in ultra-
2 cm (perf weak sandstones
D)
6 cm (perf
D)
Van den Hoek et al. Hollow Single Friable 4500 6.83 breakouts mechanical N/A Study cavity failure
(1996) cylinder phase oil Castlegate and 21C 0.18 failure with different size
16 cm (H) weakly tensile and its role on sand
7.8 cm (D) consolidated failure production
0.95–1.3 cm Saltwash South
(perf D) sandstones
Cook et al. (1994) Hollow Single Synthetic weak 1500 N/A N/A seepage 2 Changing either
Cylinder phase sandstone 21C N/A failure stress at constant
15 cm (H) Light oil mechanical flow rate, or flow
5 cm (D) water failure rate at constant
1 cm (perf stress, and
D) monitoring the
2 cm (perf L) resulting sand
production.
Alsiny et al. (1992) Hollow Single Ottawa sand 150 0.1755 Shear band seepage 15 Visual observations,
Cylinder phase oil 21C 0.17 localization failure study cavity shape
15 cm (H) mechanical and development of
15 cm (D) failure localized and
1.5 cm (perf deformation modes
D)
Kooijman et al. Cubic Oil Castlegatet 1000 Other rock softened seepage 60 Large-scale
(1992) 70 × 70 × Brine sandstone 21C strength data (failed failure experimental test.
81 cm dilated) mechanical Wellbore simulation
10.2 cm permeable failure with multiple
(hole D) material perforations, using a
With natural outcrop
perforation rock. Study the
0.6 cm (perf effect of both
D) effective stress,
14 cm (perf drawdown and
L) water cut on sand
production.
Bratli and Risnes Water Ottawa sand 1000 N/A Arch failure seepage 500 Describing the effect
(1981) Air 21C 0.25 failure of flowing fluid and
(continued on next page)

6
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

Table 3 (continued )
Model Geometry Phase Rock P (psi) E (GPa) Failure Failure Time Goals
pattern reasons of test
T (C) Poison
(min)

Cylinder the stability of sand


38 cm (H) arch and its failure
19 cm (D)
Hall Jr and Cylinder Single Ottawa sand 500–2000 N/A Arch failure seepage N/A Determine
Harrisberger 4.4 cm (H) phase Miocene 21C N/A failure qualitatively the
(1970) 9.5 cm (D) Air Oil-wet mechanical conditions required
1.1 cm (perf Water failure to form and
D) Kerosene maintain an arch
over an opening at
much higher
stresses

Coring deep oil wells and unconsolidated rock is not a simple oper­
Table 4
ation. For this reason, synthetic sandstones are used for studies. This
Parameters of reservoir condition.
replacement leads to serious errors in sand prediction researches.
Another major error in laboratory experiments is the limitation of Offset well data Reservoir rock Rock Production
type characteristic condition
combinations of two-phase flows (crude oil, water, and brine). Most
researchers studied water flow in the porous medium. However, the • Rate and amount • Sandstone • Rock Strength • Production/
of produced sand • Limestone • Cementation • injection
water flow can the increase cohesion and form stable arches or it can
• Field history • Shaliness • Poisson’s • Gas and Water
wash cement of sandstone and reduce its strength. Also, different vis­ ratio coning
cosity characteristics induce different forces in the arch. Further, most • Permeability • Flow rate
laboratory experiments can only simulate the onset of sand production • Porosity • Viscosity
and forecast the volumetric sand production rate, and more advanced • Natural • Reservoir
fracture pressure
apparatuses is needed to simulate long term running. However, such • Depth • Well radius
tests are time-consuming and costly (Rahmati et al., 2013). Temperature • In situ • Completion type
changes in injection wells and hydrate-bearing sediments are another stresses • Drilling method
important parameter (Zhu et al., 2020). (OBDa/UBDb)
It is worth noting that all the researches on sand production exper­ a
Overbalance drilling.
iments were conducted on sandstone rocks. Despite the fact that more b
Underbalance drilling.
than 60% of the world’s oil and 40% of its gas are produced in carbonate
reservoirs, rock fragmentation in these reservoirs has been completely transit time (Δtc ) is another example of one parameter method and it
ignored (Bentley, 2002). This problem is more severe in the mature says that no sand control is required below a certain depth (Tixier et al.,
carbonate oil and gas fields of the Middle East. 1975). These one parameter methods are highly dependent and do not
According to Table 3, the experimental studies show that these tests explicitly consider the reservoir pressure depletion (Δpde ) and the
usually have several hypotheses for simplification. It can be said that drawdown pressure (Δpde ). These parameters are included in the two
rock strength, fluid flow characteristics, test duration and sample di­ parameter methods. The relation between the cut-off drawdown pres­
mensions are the most important parameters in sand production tests. sure to the dynamic shear modulus (G) in another two-parameter
Researchers usually focus on only one or two of these parameters, which method was described by (Stein and Hilchie, 1972).
may lead to many errors in contrast to reality. Multi-parameter correlations can increase the prediction accuracy of
sand production. A wide range of parameters including depth, Δtc , oil
1.2. Field data production rate, Δpde , productivity index, water cut and gas cut and
shaliness. Due to of the extra data requirements the multi-parameter
Sand production prediction methods based on field data are based on methods are not widely used in the petroleum industry (Ghalambor
creating a correlation between operational parameters, sand production et al., 1989). Properties such as permeability, porosity, viscosity, rock
data of well and field. There are several parameters that can affect sand strength, Poisson’s ratio and in situ stresses, etc. Are uncertainties
production. Operationally, not all parameters are recorded because it is (Acock et al., 2004). Some uncertainty variables can be changed during
difficult in practice to monitor and record several years (Veeken et al., oil and gas production. Other parameters such as wellbore radius, and
1991). In general, there are three types of field models for predicting completion type are the result of design choices. Some parameters such
sand production: 1. Core based, 2. Sonic-log based, and 3. Regional as skin, are an uncertainty. Nevertheless, the range of skin varies
statistic models. Core based models are numerical models with curves of depending on the type of completion (Wehunt, 2003).
stress/strain as input data for calculating perforation stability strength. Generally, there are uncertainties in all input parameters obtained
Sonic based models are direct estimation from standard logs which from the field for sand production prediction tools (Wu et al., 2010).
predict cavity stability with a correlation between linear stress/strain Rock strength (UCS) has the greatest influence on well pressure pre­
and G vs. cohesive-strength. Regional Statistical models are based on diction to control the onset of sanding, which was studied by (Ogun­
field data and it is mostly used for after sand production and field kunle et al., 2018). The developed model, which is based on numerical
development (Morita, 1994). simulation and analytical solution is history matched with field data. If
The use of only one parameter is the simplest form of sand prediction there are significant discrepancies, the input parameters are
tool based on field data. The cut-off depth criterion for the sand control re-examined and corrected to fine tune the predictions (Burton et al.,
installation is a good parameter in deltaic environments (do not install 1998). Table 4 shows the parameters of reservoir condition that are
the sand control below a certain depth). The compressional sonic wave different in each field and well. The large number of these parameters

7
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

Table 5 Table 6
The most important fundamental deficiencies of numerical models related to The most important sources of inherent uncertainty in a geomechanical model.
sand production. Source of uncertainty Types
Deficiencies Reasons
Constitutive model (Gazzola Elastic model group Plastic model group
Absence of critical-state-based Researchers have tended to avoid them et al., 2019) 1. Elastic, isotropic 1. Drucker-Prager
constitutive models due, perhaps, to several calibration 2. Elastic, orthotropic model
parameters such as hardening and model 2. Mohr-Coulomb
softening 3. Elastic, model
Erosion sanding criteria are more Different natures of sand production transversely 3. Ubiquitous-joint
applicable to weak rocks and for isotropic model model
strong rocks, other sanding criteria 4. Strain-hardening/
such as tensile or shear failure are softening model
required 5. Bilinear strain-
Cavity shape hardening/
Permeability around cavity softening
Mesh size issues Simulation of perforated wellbores 6. Double-yield
requires a very fine mesh around the model
perforations, which makes the analysis 8. Modified Cam-
computationally demanding especially is clay model
3D modeling. 2D models of the 9. Hoek-Brown
perforation using plane strain and model
axisymmetric assumptions do not fully 10. Modified Hoek-
capture the accurate mechanical and Brown model
fluid flow responses. Element generation • Element shape and size
Lack of methods to capture sand arching The complexity of sand arching and its • Based on solution algorithm: polyhedral,
implementation in numerical models polygonal etc.
Lack of sufficient research for Difficult and time consuming procedure Initial conditions and Model configuration, downscaling etc.
calibration procedures in the DEM boundaries
model to determine micro material Coupling procedure (fluid • Decoupled
parameters pressure, thermal and • Explicitly coupled
mechanical stress) and • Iteratively coupled
undrained pressure response ( • Fully coupled (Dean et al., 2006)
together with their uncertainty may have a negative impact on the su­ Taron et al., 2009)
Solution method • Finite element, finite difference
periority of this method. Some researchers such as Mahmud et al. (2020)
• Continuous and discrete numerical solution
provided a smart control framework for risk mitigation by studying all • Explicit, implicit
influencing factors related to reservoir such as in situ stresses, formation
grain consistency, and the completion strategies. In some methods of
prediction, sand production data from an oil field or a single well were these simulation models are continuum methods, while a few re­
simulated with numerical models to investigate the influence of in-situ searchers have developed a method to predict sand production based on
stress anisotropy, completion type, perforation orientation, and well a discrete element model. Some numerical simulations are able to assess
placement (Alquwizani, 2013). the conditions leading to onset of sand production, while other methods
are able to predict the rate and volumetric production. Despite the
progress made in the last decades, sand production methods are still not
1.3. Numerical simulation able to estimate sand mass and its rate in a reliable form. In summary,
the most important fundamental deficiencies of numerical models are
To date, the most important approach to predict sand production has classified in Table 5.
been numerical methods. The analytical correlations and experimental Generally, numerical models contain some parameters to define the
results can help these methods to work more efficiently. However, the system. There are an infinite number of unknowns described by a finite
numerical models have their own limitations. number of model parameters. By specifying the model parameters, nu­
The mechanical and hydro mechanical instability (originated from merical models are implemented to characterize the system. These
flow-induced pressure gradient) adjacent to the wellbore are two main models can be used to predict the past and future behavior of the system.
mechanisms in numerical simulation methods. Numerical methods used These methods are known as the forward problem. The inverse problem
to simulate these mechanisms are categorized into continuum and dis­ method consists of deriving the unknown model parameters based on
continuum approaches. In the continuum approach, the elements are the past system behavior (measurements). Commonly, these compli­
considered as continuous matter when deriving the governing differ­ cated numerical models lead to a strong nonlinear calibration between
ential equations. The continuity assumption means that the material the parameters of the model and its prediction or simulation of the data
components cannot be decomposed into smaller parts (Jing and Ste­ (Trampert, 1998). This strong nonlinear calibration can also be
phansson, 2007). On the other hand, the discrete element method (DEM) considered as one of the shortcomings of prediction methods.
simulates sand production with the modeling of bulk behavior of gran­ In most calibration models, there are two types of errors, namely
ular materials, especially to discover the mechanism of sand production. measurement errors and model errors. Measurement errors are defined
The large computation time is a problem for large scale simulation. at the time of measurement and usually have known statistics. These
Furthermore, it is difficult to calibrate these models because there are errors result in uncertain petroleum reservoir data. Field information is
various uncertainties associated with creating a model with the exact generally noisy and sparse. These data are usually obtained from finite
particle arrangement that resemble the real material. In addition, cores (a very small portion of the reservoir). The data can also be time-
methods to directly measure the sandstone micro properties have not yet dependent over large scales of the reservoir (Subbey et al., 2004).
been established (Belheine et al., 2009; Kulatilake et al., 2001). There­ Model errors arise from resolution (including mesh number, mesh
fore, continuum models are more suitable for field-scale problems. size etc.), estimation, and lack of comprehensive modeling of all of the
However, there are hybrid models that use continuum and discontinuum relevant physics. Furthermore, model prediction accuracy is largely
methods. dominated by uncertainties stemming from the definition of alternative
Rahmati et al. (2013) and Subbiah, 2020 have fully reviewed the conceptual models and their characteristics. The main sources of
numerical models in a review article. Based on their reviews, most of

8
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

Fig. 2. Sources of error in sand-production modeling from physical model to solution (Morita, 1994).

Fig. 3. The experimental setup units and physical modeling view of sand production. After Fattahpour et al. (2012).

inherent uncertainties in the model are categorized in Table 6. images of the hole after failure. The setup of these tests consists of a
The evaluation of modeling errors is a very difficult task and there­ triaxial cell, hollow cylinder samples, sand detection and measurement
fore they have been mostly ignored in the literature (Sambridge and apparatus. The units of experimental setup and the physical modeling
Mosegaard, 2002). For example, in petroleum reservoir simulation, view of sand production are shown in Fig. 3.
many efforts have been made to correct the modeling error due to the The hollow cylinder synthetic sample and different parts of the
use of coarse grids (Christie et al., 2002; James et al., 2001). The cor­ triaxial cell are shown in Fig. 4. Other parts of the experimental setup
rectness of the model depends on each of these errors. If the model errors include the fluid tank, the fluid injection pump, the sand measurement
are trivial, when the model is calibrated by measured data, there is some device, the axial and radial stress pump, the data acquisition device, and
degree of reasonable predictability. If there is a small error in the the computer. The confining stress and fluid flow are applied to the
modeling, the results of the model are not reliable and usually lead to samples axially and radially during the test. The test process will be
incorrect management decisions (Carter et al., 2006). Fig. 2 simply briefly explained later and the detailed procedure of the experimental
shows error sources in sand-production modeling from physical model test can be found in Fattahpour et al. (2012).
to solution. The synthetic samples were prepared with proper particle size dis­
tribution (PSD) and mixed with cement and water. Using PSD method,
2. Experimental test and numerical simulation the average grain size diameter is calculated by evaluating the formation
sand samples (Fattahpour et al., 2012). The particle size distribution of
In this study, a hollow cylinder sand production experiment and a the synthetic sandstone is similar to that of a natural sample from an
simple widely used numerical simulation are presented to show the sand Iranian oil reservoir (Fig. 5).
production numerical model’ deficiencies. Since most studies use con­ The mechanical properties of the synthetic sandstone with 15 cm in
tinuum approaches, it was decided to use a simple finite difference diameter and 30 cm in height are shown in Table 7.
program whose governing equations include the continuity, momentum, To perform the numerical modeling, some necessary parameters
and Darcy equations. In addition, a number of methods were used to including shear modulus, bulk modulus, tensile strength, and cohesion
reduce undesirable influencing factors such as mesh size. are calculated using the following formulas and relationships according
to the characteristics of sandstone (Van den Hoek and Geilikman, 2003;
2.1. Experimental setups (hollow cylinder sand production experiment) Zeng et al., 2016). These parameters are shown in Table 8.
E
Generally, a hollow cylinder core is used to simulate the sand pro­ G= (1)
2(1 + v)
duction (Veeken et al., 1991; Vriezen et al., 1975). The results of these
tests are a kind of graph (rate and cumulative produced sands) and

9
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

Fig. 4. Triaxial cell parts designed for hollow cylinder synthetic sample tests. After Fattahpour et al. (2012).

Table 8
Calculated shear modulus, bulk modulus, cohesion, and tensile strength.
Cohesion (MPa) Bulk Modulus (K) Shear modulus (G)
(GPa) (GPa)

3.426 1.572 1.279

E
K= (2)
3(1 − 2v)

UCS
C= ( ( )) (3)
2 tan 45 + ϕ2

Where E, v, are the Young’s modulus and Poisson’s ratio of the rock. UCS
and ϕ are the uniaxial compressive strength and friction angle.

Fig. 5. PSD of the sand mix is similar to a natural cores from an Iranian oil
2.1.1. Experimental procedure
reservoir. After Fattahpour et al. (2012).
During the test, confining stress and fluid flow are applied to the
samples radially and axially. The radial stress is gradually increased and
fluid is injected at different fluid flow rates at each stress step. The axial

Table 7
Mechanical properties of synthetic sandstone. After Fattahpour et al. (2012).
Sample Biot Friction angle Density Cement Porosity Permeability UCS Poisson young modulus Tensile strength
coefficient (Degree) (kg/m3) content (%) (%) (mD) (MPa) ratio (GPa) (MPa)

D20 0.7 30 2300 20 24 210 11.87 0.18 3.02 1.1187

10
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

2.2. Process of numerical analysis

In this study, the numerical program uses an explicit or implicit,


time-marching Lagrangian analysis (Cundall and Board, 1988). Me­
chanics of the medium are derived from general principles of strain
definition, laws of motion, and the constitutive mathematical equations
determining material behavior. The variables of mechanics (stress) and
kinematics (strain rate, velocity) are related together with the derived
mathematical expression (a series of partial differential equations).
These equations are solved for given properties, geometries, specific
boundary and initial conditions. The approach of solution is defined by
the finite difference, i.e., the first-order space and time derivatives of a
variable are estimated by finite differences, considering linear variable
variations in a finite space and time interval, respectively.
The formulation of the coupled processes of mechanics and hydro­
mechanics is carried out in the scheme of quasi-equilibrium Biot theory,
together with the single phase Darcy flow in a porous medium. The
variables of fluid flow in porous media are saturation, pore pressure and
specific flow vector.
The solving method of geomechanics and fluid flow equations can be
categorized into sequential and fully coupled approaches. The flow and
geomechanics unknowns are calculated concurrently for one time step
Fig. 6. Applied stresses and fluid boundaries. (fully coupled) or sequentially by detaching the coupled problem into
sub problems (White et al., 2016). One of the straightforward sequential
stress is considered to be one-third of the radial stress, in lack of any field methods is one-way coupling. In this method, the fluid-flow sub-prob­
stress measurement data, based on a general assumption that the lem is solved earlier and then the effect on the geomechanics properties
lithostatic-horizontal stress is equal to one-third of the axial stress. is defined in the next step (Beck et al., 2020). In this study, calculation of
Radial stress and axial stress represent the vertical and horizontal stress, the flow equation is performed in parallel with the mechanical model
respectively (Fattahpour et al., 2012). (fully coupled) to implement the effects of the fluid-solid interaction, the
The change in radial stress and fluid flow simulates the well pressure relationship between the mechanical volume alterations and the pore
reduction or reservoir depletion. This depletion could cause an increase pressure change.
in effective stress. It is assumed that the resulting excess stress is the In this research, the finite difference numerical program uses an
major cause of sand production. For this test, diesel fuel is used as the explicit and implicit time-marching Lagrangian analysis. Although the
fluid because live oil viscosity in the depth of reservoir is close to the implicit approach is second-order accurate, it is needed to choose the
diesel fuel viscosity. The density and viscosity of the injected fluid are proper time step. Technically, this value must be lower than the wave­
0.82 g/cm3 and 4.5 cP, respectively. Stresses and fluid boundaries are length of any nodal pore-pressure variation. The explicit approach is
applied to the sample as shown in Fig. 6. implemented earlier in the modeling or in its perturbed phases. Then the
The stresses are incrementally applied to the sample according to implicit method is utilized for the rest of the simulation. By preference,
Fig. 6 until the sands were produced at 27 MPa radial stress. The pro­ the implicit approach can be used with the explicit time step value for
cedure and amount of the injected fluid into the cell is also shown in higher accuracy (Pyatigorets and Russell, 2020).
Fig. 7. In the implicit approach, a set of equations which requires at least
three iterations must be calculated at each time step. The amount of
calculation needed for one iteration step is nearly equal to that required

Fig. 7. Stresses and fluid applied stepwise to the sample for 27 MPa radial stress and 9 MPa axial stress.

11
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

Fig. 10. Friction hardening and cohesion softening relations as a function of


plastic shear strain (Vermeer and De Borst, 1984).

According to the Mohr–Coulomb elastic-perfectly plastic model, shear


and tensile failures occur when the stress tensor meets the tension cutoff
and shear envelope, respectively.
Permeability and porosity changes during testing can be updated as a
function of current volumetric strains using the following equations
(Nouri et al., 2004).
Fig. 8. Mechanical and hydrolytic cycle scheme. Δ∅
εv = (4)
1− ∅

Where εv is the volumetric strains. Then, the variation of permeability


was estimated by applying the equation Kozeny–Carman (Das, 2013).

C ∅3
k= (5)
γf (1 − ∅)2

Where ∅, γ f , c are the porosity, the specific weight of the fluid, a constant
which is depended on the specific gravity, the permeability and the
porosity.
In the numerical algorithm, the initial pore pressure and the internal
hole pressure were set to ambient pressure and full saturation. The fluid
pressure and stresses were gradually applied based on experiment into
the external boundaries. According to the nature of sand production
physics, a coupled fluid-mechanical calculation was performed. After
each step, a check was conducted through the entire model to find the
elements that met the sanding criteria. Moreover, an additional criterion
was implemented to eliminate eligible elements from the hole surface,
and then their volume was added to the cumulative sanding volume.
Concurrently, the boundary conditions were reset at the new boundaries
inside the hole. Fig. 8 shows a schematic representation of the model
cycles. At each step, the simulations were iterated between mechanical
and hydrolytic cycles.
In this study, the model was generated by a cylindrical shaped mesh
(consisting of polyhedral or polygonal elements). Because of the sym­
metry of the model, a quarter of the model was modeled to decrease run-
time (Fig. 9). In addition, finely spaced elements are used near the hole
(zone of interest) with pattern of increasing the size of the elements
radially. Finer meshes lead to more precise result, leading to a better
representation of high-stress gradients.
The numerical modeling was developed based on the two Mohr-
Fig. 9. Hollow cylindrical model constructed with polyhedral shaped mesh.
Coulomb elastic-perfectly plastic and Mohr-Coulomb cohesion soft­
ening/friction hardening failure criteria. In the Mohr-Coulomb elastic-
for one time step in the explicit method. Further, runtime and computer perfectly plastic the defined parameters of shear modulus, bulk
memory are two important parameters that should be considered when modulus, cohesion, friction angle and tensile strength are implemented
choosing the implicit method. This method is applied in FLAC3D. This in the numerical model. The input parameters of strain cohesion soft­
software is used here because it facilitates the user to write necessary ening/friction hardening failure criterion based on mobilized mechan­
code to dynamically change the model properties during numerical ical properties include shear modulus, bulk modulus, cohesion, friction
simulation based on the model state, and also the conceptual model is angle and tensile strength.
originally able to transfer to other software packages. Based on various levels of hardening parameter, the new magnitudes
In this simulation, the Mohr-Coulomb elastic-perfectly plastic and of friction angle (ϕ* ), and cohesion (c* ) were estimated by Eqs. (6) and
Mohr-Coulomb cohesion softening/friction hardening failure criteria (7). The relations between friction hardening and cohesion softening as
were used. Axial symmetry was considered in the simulations.

12
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

Table 9
Mobilized mechanical parameters.
Shear hardening parameter % 0 0.3 3 5 7 10 20

Φ 0 9.68 24.9 28.12 29.4 30 30


C 20 × 106 9.2813 × 106 6.4987 × 106 3.4262 × 106 1.312 × 106 0.1706 × 106 1.048

[ ( )2 ]
Table 10 ε− p εf
Well, casing and perforation data (Papamichos and Malmanger, 1999). c* = c exp − (7)
εf
Perforation intervals between 2358.4–2477.3 m TVD
Total perforated length of the well 71 m Fig. 10 shows that the mobilized friction angle gradually increases
Well inclination angle with vertical 50◦ with plastic shear strain to reach the limit friction angle (ε-p reaches the
Perforation density 6 SPF
constant εf). Likewise, the variation of cohesion is controlled by plastic
Perforation phasing 60◦
Number of perforations 1400 shear strain with a limit consideration (εf).
Number of productive perforations 700 The mobilized mechanical parameters used in the study are outlined
Perforation length 0.286 m in Table 10. In this table, friction angle, and cohesion, are calculated as a
Perforation radius 0.01 m function of shear hardening parameter (accumulated plastic shear
Casing Young’s modulus (E) 250 GPa
strain).

a function of plastic shear strain are schematically shown in Fig. 10. 2.2.1. Fracture energy regularization
Experiments show that these curves can generally be simply expressed In the hardening phase, the rock deforms almost uniformly. There­
by the following friction angle and cohesion functions (Vermeer and De fore, the whole sample deformations are uniform and independent of the
Borst, 1984). mesh shape and size. In the softening phase, on the other hand, the
√̅̅̅̅̅̅̅̅̅̅̅ deformation is localized in the shear bands. During numerical solving,
sin ϕ* = 2
ε− p εf
sin ϕ for ε− p < εf (6) the shear band splits into the smallest possible thickness up to the size of
ε− p + εf one row of elements (Bažant et al., 1993). The shear band size affects the
energy amount released for this localized deformation. The larger ele­
sin ϕ* = sin ϕ for ε− p > εf ments have a higher energy release rate. The difference of released en­
And, ergy amount for mesh sizes leads to mesh-dependency. In the hardening
phase, the mesh dependency does not occur as long as the entire sample
deforms uniformly and the equal energy is applied to deform the sample
with different mesh sizes (Crook et al., 2003; Pietruszczak and Mroz,

Fig. 11. Mesh dependency observed in softening regime, smaller mesh size results in more realistic modeling (left), effectiveness of fracture regularization method
and resolve mesh dependency (right) (Jafarpour et al., 2012).

Fig. 12. Objective dissipation via fracture energy theory (Crook et al., 2003).

13
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

1981). Table 11
Jafarpour et al. (2012) examined the effectiveness of this method in Reservoir data (Papamichos and Malmanger, 1999).
sand production prediction by simulating regular coarse and fine mesh. Porosity 0.177–0.267_ average = 0.22
In Fig. 11 (left) illustrates the mesh dependency observed in the soft­ Permeability 90-4004_ average = 400 mD
ening regime. Smaller mesh size results in more realistic modeling. Initial reservoir pressure 38 MPa
Current reservoir pressure 29 MPa
Fig. 11 (right) displays the effectiveness of the fracture regularization
Fluid viscosity 5 cP
method and resolves the mesh dependency. The element size variations Density of fluid 0.84 × 103 kg/m3
have little effect on the numerical result. Total vertical stress 49.1 MPa
The localization of plastic strain is supposed to occur in the finite Total minor horizontal stress 44.5 MPa
volume of the crack band (Fig. 12), which provides finite energy Total major horizontal stress 45.8 MPa
dissipation. Young’s modulus (E) 2.66 GPa
Poisson’s ratio 0.24
There are several approaches to regularize continuum numerical
Density of sand particles 2.65 × 103 kg/m3
modeling to reduce mesh dependency. Cosserat continuum, gradient
plasticity models, non-local continuum, and fracture energy regulari­
zation are the most commonly used methods. The Cosserat continuum
improves the standard continuum by changing the standard kinematic
variables with micro-rotations (Mühlhaus and Vardoulakis, 1987).
Papanastasiou and Vardoulakis (1992) demonstrated that the
micro-rotations have a significant effect on the borehole stability com­
putations by applying the Cosserat continuum model. However, a po­
tential shortcoming of this method is that it is only practical as a
regularization method when frictional slip is dominant in rock failure
(Crook et al., 2003). The essential feature of the gradient plasticity
method is that the behavior of the yield function is related to the spatial
gradients of the damage variable in addition to the damage variable. At
the microstructural level, these higher order gradients in a macroscopic
constitutive model are correlated with non-local interaction. Similar to
the Cosserat method, the gradient plasticity method is appropriate
method in finite strains and complex material models (Mühlhaus and
Vardoulakis, 1987; Pamin, 1994). To study elastic randomly heteroge­
neous microstructure materials, the non-local method was developed.
This method considers a relation between non-local spatial means of Fig. 14. The data of sand rate and the applied drawdown over a 120 h. The first
20 h of data are used in numerical simulation, with a constant drawdown
both continuum stress and strain over a given volume (Eringen, 1966).
pressure of 2.5 MPa due to the run time constraints (Papamichos and Mal­
The non-local method can be computationally problematic in field study
manger, 1999).
(Crook et al., 2003).
In this study, fracture energy regularization is applied. The advan­
tage of this approach is the ease of implementation in any finite strain Where ϵf (e) and ϵf are the modified hardening parameter and hardening
numerical model and different constitutive models. This method in­ parameter of material, respectively. In the softening region, Eq. (9)
cludes a material characteristic length, hc, and a characteristic length of equalizes the energy release (area under the stress–strain curve, Fig. 12)
element, h(e). The finest applicable mesh is when hc = h(e), where h(e) for different element sizes.
defined as the diameter of the sphere having the equal volume to the
element under consideration. The characteristic length of the element 2.2.2. Material characteristic length
considering the element volume can be calculated using the following The material characteristic length, hc, is considered to be equivalent
equation: to the shear band thickness based on literatures. Many research pro­
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ posed that the shear band thickness (ts), is about 10–20 times the mean
h(e) = 2
3 3(vol.ele)
(8) grain size (d50) of the material (Desrues and Hammad, 1989; Yoshida,
4π 1995). They concluded that most of the data are coincided with the line
To use larger element sizes, it is necessary to alter the hardening ts = 10 × d50.
parameter such that it causes the same fracture energy dissipation. The However, the equivalence between material characteristic length
hardening parameter is changed proportionally to element sizes such and the shear band thickness is true as long as the shear band occurs in
that it results the same fracture energy (Crook et al., 2003). only one row of elements. Numerical tests indicated that when the ori­
ented mesh angle is the same as that of the failure angle, the shear band
hc f will be constrained to one element row and the material characteristic
εf (e) = ε (9)
h(e)

Fig. 13. Procedure of implementing the fracture energy regularization method in the numerical model.

14
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

length will be the same as the shear band thickness. In this study, the sand production numerical model and three-dimensional stress trans­
orientation of mesh is not the same as that of the failure angle of the formation is done to calculate the model boundary conditions appro­
shear band. priate for inclined wells and oriented perforations.
Dependency analysis of the numerically created shear band thickness A reference perforation hole adjusted to the positive y-direction is
on the material characteristic length for a quadrilateral element based modeled in the wellbore. The following equation is applied to convert
on Nouri et al. (2009) procedure, shows that the shear band for quad­ the stresses from the coordinate system (x, y, z) to a new coordinate
rilateral shaped elements would be approximately 3.322 times the system (x′ , y′ , z′ ). More detailed information can be found in the works
characteristic length. Thus, the characteristic length of material is of (Alquwizani, 2013; Rösler et al., 2007)
considered to be about 3.01 times the mean grain size
(11)

σ ij = α.σ ij .αT
hc = 10 × ​ d50 3.322 = 3.01 × d50
/
(10)
⎡ ⎤ ⎡ ⎤
σ 11 σ 12 σ13 σH 0 0
The mean grain size of the sample is 0.00708 in. (0.18 mm). Thus, ⎣ σ 21 σ 22 σ23 ⎦ = α.⎣ 0 σh 0 ⎦ .αT (12)
the characteristic length of the material would be nearly 0.0213 in.
σ 31 σ 32 σ33 0 0 σv
(0.5418 mm).
Fig. 13 shows the procedure of implementing the fracture energy
here σij is the stress tensor in the new coordinate system and α is the

regularization method in FLAC3D by FISH programming language.


orthogonal transformation matrix. These values are the direction cosines
In each step, ϵf(e) is used to obtain the ϕ* and c*. New ϕ and c are
between the new and old coordinate systems. The elements of the
updated in Table 9 corresponding to the value of the shear hardening
transformation matrix, α, can be calculated using Euler angles instead of
parameter. In other words, a specific behavioral model is created for
using nine angles. The superscript T represents the transpose of the
each element.
matrix.
Using Euler angles, the y-x-z rotation is applied by the following
2.3. Field data angles.
π
(13)
′ ′
The data of a real oil well and numerical simulations of sand pro­ ψ =β −
2
duction in a North Sea reservoir are provided to examine the applica­
β : Wellbore azimuth from σH direction.

bility of prediction numerical model. Based on the literature, most field
studies have generally been conducted to estimate the onset of sand ψ : The rotation angle around the original z-axis.

production. The current study of well data and numerical model at­ The corresponding rotation matrix is
tempts to establish the relationship between sand production rate as a ⎡ ′ ′ ⎤
cos ψ sin ψ 0
function of time, stress, and flow rate. These analyzes help the field
α1 = ⎣ − sin ψ cosΨ 0 ⎦ (14)

management and well production strategy to produce with maximum


0 0 1
productivity and minimum sanding problems (Papamichos and Mal­
manger, 1999). θ : The inclination of the borehole, measured from the vertical. The

Sand production rates were measured continuously at an oil well in a corresponding rotation matrix is
North Sea reservoir. Table 10 shows the perforation data, such as total ⎡ ⎤
1 0 0
perforation length, phasing, and perforation density. The total number
(15)
′ ′
α2 = ⎣ 0 cos θ ′ sin θ ′ ⎦
of perforations is about 1400 and it is considered that 700 perforations
0 − sin θ cos θ
are actually not producing.
The well inclination is 50◦ in the perforated section. In addition, the ϕ : Orientation angle of the reference perforation, measured coun­

reservoir data, such as porosity and permeability, in situ stresses, initial terclockwise from high side of the hole. The corresponding rotation
reservoir pressure are presented in Table 11. Fig. 14 illustrates the data matrix is
of sand rate and the applied drawdown over a 120 h. However, for the ⎡ ′ ′ ⎤
numerical simulation, the first 20 h of data are used with a constant cos ϕ sin ϕ 0
α3 = ⎣ − sin ϕ cos ϕ 0 ⎦ (16)
′ ′

drawdown pressure of 2.5 MPa due to the run time constraints. (Papa­
michos and Malmanger, 1999). 0 0 1
The borehole inclination at the perforated interval is 50◦ . To simu­ The matrix, α can be calculated based on these rotation angles as
late this inclined well, the in-situ stresses of the field are inserted into the

Fig. 15. Generated meshes to simulate wellbore and perforations.

15
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

Fig. 16. Wellbore (8 ½ in) and casing (6 5/8 in) numerical model with 6 perforations and 60◦ phasing.

Fig. 17. Experimental data of sand production rate as a function of time, for 27, 16.5, 15 and 12 MPa radial stress, in sample with two different perforation hole sizes
of 10 and 20 mm. After Fattahpour et al. (2012).

follow modeled in a cube with a size of 3 × 3 × 3 m (Figs. 15–16). The modeling


is done with same sized element and smaller mesh size is used around
α = α1 . α2 . α3 (17)
wellbore. The amount produced sands of 6 perforations is multiplied by
Similarly, the gravitational acceleration vector, gi can also be

116 to simulate 700 perforations. Hole and casing diameters are 8 ½ and
transformed according to the following formula 6 5/8 inches, respectively. The Young’s modulus (E) of the casing are
assumed to be 250 GPa.
(18)

gi = α⋅gi
⎡ ⎤ 3. Results and discussion
0

gi = 0 ⎦
In order to scrutinize the shortcomings of numerical models and
− 9.81
experimental tests a simple numerical method was performed and
The 3D model of a wellbore with 6 SPF and a phasing of 60◦ is compared with laboratory data. The results of sand production tests,

16
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

Fig. 18. Experimental sand production rate (black line) and numerical simulation with Mohr-Coulomb elastic-perfectly plastic (A) (blue line), numerical simulation
with strain cohesion softening/friction hardening (B) (blue line), the red line shows the trend line of the numerical simulation. (For interpretation of the references to
colour in this figure legend, the reader is referred to the Web version of this article.)

Fig. 19. Experimental sand production cumulative (black line) and numerical simulation with Mohr-Coulomb elastic-perfectly plastic (A) (blue line), numerical
simulation with strain cohesion softening/friction hardening (B) (blue line), the red line shows the trend line of numerical simulation. (For interpretation of the
references to colour in this figure legend, the reader is referred to the Web version of this article.)

Fig. 20. Cross-section of the experimental sample and three stages of sand production (A). After Fattahpour et al. (2012).– Cross-section of numerical simulation by
Mohr-Coulomb elastic-perfectly plastic model (B) - numerical simulation by strain cohesion softening/friction hardening model (C).

17
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

Fig. 21. Bottom view of the cavity in the cylinder sample after testing (A). After Fattahpour et al. (2012). - Bottom view of numerical model in Mohr-Coulomb
elastic-perfectly plastic (B) - Mohr-Coulomb cohesion softening/friction hardening failure criteria model (C).

element size which increases radially from the inner radius to the outer
radius. In the early sand production, the recorded rate is related to the
small element size at the cavity wall and with the progress of producing
the larger element size affects the rate of sand production. This is one of
the disadvantages of continuous numerical modeling. The discrete nu­
merical modeling with the same size element can solve this problem.
In contrast to the sand production rate results, the cumulative nu­
merical results for both models agree relatively well with the experi­
mental cumulative data. As can be seen from Fig. 19 A-B, the slope in the
early cumulative sand production is small due to the difference in
element and increases as production progresses.
Fig. 20A shows the cross-section of a sample. In experimental
observation, there are three stages of sand production. In the first stage,

Fig. 22. Supposed that sand slabs detach from hole surface when localization
occurs and accumulative shear plastic reaches 0.1.

fluid rate steps and sand production rate during time, for 27, 16.5, 15
and 12 MPa radial stress, in sample with two different perforation hole
sizes of 10 and 20 mm are shown in Fig. 17.
From the experimental data and literature review it is clear that
under constant conditions of rock type, fluid properties and time,
changes in stress value, fluid rate and perforation diameter have a great
effect on the sand production trends.
Figs. 18–19 compare the rate and cumulative sand production
experimental results (test data of 27 MPa radial stress, 9 MPa axial stress
and perforation diameter of 20 mm) with the Mohr-Coulomb elastic-
perfectly plastic and Mohr-Coulomb cohesion softening/friction hard­
ening model, respectively. As can be seen in Fig. 18 A-B, the graph of
Fig. 23. Different element size near the cavity in a given model with the radial
sand production rate (blue line) does not have acceptable match to the
increased elements (A) - Same size elements in the structure of the numerical
experimental data (black line). This difference originates from different
model (B).

18
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

Fig. 24. Different element sizes in Mohr-Coulomb elastic-perfectly plastic (radial increased elements pattern). Elements are reduced from A to D. Experimental
data (black line) - Numerical simulation (blue line). The number and radial dimensions of the elements are specified at the top of each shape. (For interpretation of
the references to colour in this figure legend, the reader is referred to the Web version of this article.)

Fig. 25. Different element sizes in Mohr-Coulomb cohesion softening/friction hardening failure criteria model (radial increased elements pattern). Elements
are reduced from A to D. Experimental data (black line) - Numerical simulation (blue line). The number and radial dimensions of the elements are specified at the top
of each shape. (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)

19
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

a small cavity is formed at the dead end. Produced sands from borehole 3.1. Sensitivity analysis on the size and shape of elements
wall create a conical shape with a slight slope in the second stage and
third stage occurs during the second conical shape creation. The nu­ In order to show the effect of element size and shape on the nu­
merical modeling shows a relatively similar trend compared to the merical simulation results a series of sensitivity analysis was performed
experimental stages in the both Mohr-Coulomb elastic-perfectly plastic in the presence of the regularization method to reduce the effect of the
and Mohr-Coulomb cohesion softening/friction hardening failure different released energy amount of each mesh size. In other words,
criteria models (Fig. 20B–C). larger elements are stiffer than smaller ones and this method corrects
Fig. 21A shows the bottom view of the sample after testing. The these differences as previously described. However, in a specific model,
produced sands are separated from the hole surface due to shear bands the mesh sizes increase radially. This means that the separation of the
in the form of curved slabs. The bottom view of the numerical model also larger element far from the cavity has a larger volume of sand particles
shows the acceptable similarity with the experimental test in the Mohr- than the smaller element, and it makes the results not being similar to
Coulomb elastic-perfectly plastic and Mohr-Coulomb cohesion soft­ each other in different element sizes. Fig. 23 illustrates different element
ening/friction hardening failure criteria model (Fig. 21B–C). sizes near the cavity (radial increased elements pattern) in a given model
As can be seen, the numerical modeling correctly shows the onset of and the same size elements in the structure of the numerical model.
sand production and also the formation failure stages, but is extremely Figs. 24–25 show the sensitivity analysis on the size of the elements
weak in terms of production rate of sands. This error is mainly due to the (radial increased elements pattern) in the Mohr-Coulomb elastic-
size and number of meshes as well as the sand production criteria. perfectly plastic and Mohr-Coulomb cohesion softening/friction hard­
Sanding occurs only from the elements of the cavity face when the ening failure criteria model, respectively. These different element sizes
accumulative shear plastic reaches 0.1. It was observed in the laboratory make the initial part of the results underestimated and also leads to
that the rock around the hollow is broken into pieces similar to the shear overestimation at the end of the modeling.
bands in the numerical modeling when the accumulative shear plastic Figs. 26 and 27 illustrate sensitivity analysis for the size of the ele­
reaches 0.1. Also, the effect of cement wash on rock failure was not ments in the Mohr-Coulomb elastic-perfectly plastic and Mohr-Coulomb
observed. It is very obvious that the longer time needed to wash the cohesion softening/friction hardening failure criteria model, respec­
cement, especially in the presence of oil fluid. Therefore, the failure tively (same sized elements pattern). As can be observed in these
happens more mechanically. When the accumulative shear plastic of an models, mesh sizes do not make much difference to the results.
element is greater than 0.1, it is removed from the model. The specified
critical amount of accumulative shear plastic, 0.1, is defined by shear
plastic localization. Sand slabs are assumed to detach from the hole 3.2. Accuracy of the models
surface when localization occurs (Fig. 22).
Removing or retaining the elements that meet the sanding criteria is To have an indicator of model performance, the cumulative proba­
an important decision in numerical sand production modeling and de­ bility distribution versus absolute relative error diagram is also used.
pends on reservoir features, experimental observations, engineering Fig. 28 displays the cumulative probability of sand production models as
judgment, and implemented sanding criteria. Two methods can be a function of absolute relative error (ARE). To calculate the cumulative
applied to remove the material: 1) element removes once it satisfies the probability for a specific variable, one needs to have the probability of
sanding criteria and, 2) using an adaptive meshing design and deleting all individual measurements. For a discrete distribution problem with a
of the elements from the cavity surface when the sanding criteria are total number of N cases the probability of the individual measurements
fulfilled (Nouri et al., 2009). For retained material, the residual stiffness is 1/N. For example, if there are 50 data points, the probability of the
and strength properties can be modified (decrease in young modulus or individual measurements is 1/50.
cohesion). Therefore, by adding the probabilities for all values less than or equal
In this study, sand slabs detached from the hole surface during to the specified value, the cumulative probability is calculated. This
localization and produced based on our experimental observation means that for a random sample of size N, the cumulative probability is
(Fig. 21 A). This condition indicates the disaggregated sand removal. the number of observations that are less than or equal to the specified
Therefore, we removed the elements when they met the sanding criteria value divided by the total number of N observations. The individual
in each step. observations probability must be between 0 and 1, and the sum of all
probabilities must be 1 (Naderi et al., 2019). Also, the absolute relative
error is calculated by subtracting the numerical data and the experi­
mental data at each point.
In Fig. 28, as it can be seen that for a given value of ARE, as the

Fig. 26. Different element sizes in Mohr-Coulomb elastic-perfectly plastic (same sized elements pattern). Elements are reduced from A to B. Experimental data
(black line) - Numerical simulation (blue line). The number and radial dimensions of the elements are specified at the top of each shape. (For interpretation of the
references to colour in this figure legend, the reader is referred to the Web version of this article.)

20
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

Fig. 27. Different element sizes in Mohr-Coulomb cohesion softening/friction hardening failure criteria model (same sized elements pattern). Elements are
reduced from A to B. Experimental data (black line) - Numerical simulation (blue line). (For interpretation of the references to colour in this figure legend, the reader
is referred to the Web version of this article.)

Fig. 30. Sand production rate of an oil well with 50◦ inclination and 700
Fig. 28. The cumulative probability of Mohr-Coulomb elastic-perfectly plastic perforations with 60◦ phasing (black line) and numerical simulation with Mohr-
and Mohr-Coulomb cohesion softening/friction hardening failure criteria model Coulomb elastic-perfectly plastic (blue line), the red line shows the trend line of
are shown as solid and dotted lines, respectively (radial increased elements the numerical simulation. (For interpretation of the references to colour in this
pattern). The mesh sizes reduce from A to D. figure legend, the reader is referred to the Web version of this article.)

Fig. 28 shows the cumulative probability of radial increased ele­


ments pattern models. It can be seen that increasing the size of the
meshes significantly reduces the modeling error in Mohr-Coulomb
elastic-perfectly plastic method. In the Mohr-Coulomb cohesion soft­
ening/friction hardening failure criteria model, the results are almost
identical. However, the mesh size still affects the magnitude of the error.
For the absolute relative errors smaller than 2 g, the cumulative prob­
ability of the Mohr-Coulomb elastic-perfectly plastic method (solid line
D) is somewhat better than the other numerical models. The analysis of
error depicts that 90% of the sand production data can be predicted with
an absolute relative error of less than 12 g, respectively using cohesion
softening/friction hardening failure and Mohr-Coulomb.
As can be seen in Fig. 29 with same sized elements pattern, Mohr-
Coulomb elastic-perfectly plastic and the Mohr-Coulomb cohesion
softening/friction hardening failure criteria models respectively have
better forecasting performance for sand production in terms of cumu­
lative probability versus the absolute relative error. However, the error
Fig. 29. Cumulative probability of Mohr-Coulomb elastic-perfectly plastic and amount in both models increased with decreasing mesh size. Figs. 28
Mohr-Coulomb cohesion softening/friction hardening model are shown as solid and 29 can provide further useful information on the performance of the
and dotted lines, respectively (same sized elements pattern). The mesh sizes
different prediction methods by comparing the values of ARE at the
reduce from A to B.
cumulative probabilities of 10% and 90% for all approaches. Obviously,
the results are less error-prone when P10 is closer to P90. Comparing all
cumulative probability raises, model prediction ability increases due to models, we can see that this difference is the smallest for same sized
a greater number of data points can be predicted with less than that of elements pattern.
the fixed absolute relative error.

21
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

Fig. 31. Progressive sand production that begins at perforation tips and extended to over a wide area (A to C).

The sand production results of the inclined well simulation with 6 SPF in North Sea Reservoir.
perforations and 60◦ phasing are shown in Fig. 30. This simulation was The review section of this study showed that experimental studies
done by sing same sized elements. As can be seen, the trends of the usually have several hypotheses for simplification. It can be said that
produced sands are in an acceptable agreement with the real sand rock strength, fluid flow characteristics, test duration and sample di­
production. mensions are the most important parameters in sand production tests.
Fig. 31 illustrates the casing and rock model after several hours of Researchers usually focus on only one or two of these parameters, which
simulation. As can be seen the area around the well with a radius of may lead to many errors in contrast to reality. Additionally, according to
about half a meter is strongly affected by sand production. After a while, numerical review there are many uncertainties and errors in numerical
the two adjacent perforations merge and change the trend of sand pro­ simulation, which are categorized into input parameters and numerical
duction (Fig. 31A–B). Figure C shows a catastrophic sand production model.
where unconsolidated rock intrudes toward the well. Although there are The results of the numerical model with comparison to the experi­
high fluctuation and uncertainties in the results, it seems this level of mental data showed that both failure criteria can correctly predict the
accuracy gives a good insight into decision making about well comple­ onset and stages of sand production. On the other hand, the rate and
tion type, sand control selection, and field management. cumulative of sand production were not in acceptable agreement with
In this study, it is worth noting that although models with same sized the laboratory results. Nevertheless, they provide appropriate data on
elements pattern have the lowest error amount, they predict an average the sand production trends.
of 1.5–2 times more or less with respect to reality at best. Moreover, care The main conclusion that could be drawn from this study using the
must be taken that drastically reducing the size of the meshes can still cumulative probability plot versus the absolute relative error analysis
causes errors in both mesh patterns. was to show how the size and pattern of the elements in the numerical
In sum, the results indicate that the differences between modeling simulation could affect the final results. The underestimation at the
and experimental results are very large for some models. Errors origi­ initial part and the overestimation at the end of results arose from the
nated from uncertainties in the numerical model and several hypotheses different element size, which increases radially from the inner radius to
for simplification in experimental test are the reasons for these dis­ the outer radius. Applying same size element greatly reduced this defect.
agreements. In this study, the results are very sensitive to the mesh size The superiority of behavioral models could not be assessed properly in
and shape in the numerical model and the stress value and fluid flow in the shadow of this deficiency. Further experimental investigations are
the experimental test. Another important issue is the absence of a needed to estimate test errors in the presence of simplifications.
comprehensive benchmark for the sand production criteria due to the Despite all the errors and uncertainties in both the laboratory and
different failure patterns and cavity shapes. There are other important modeling studies, the simulation of a real sand production in North Sea
parameters that not addressed in this study and have only been theo­ Reservoir with same sized element was relatively acceptable. Although
retically reviewed. there is high fluctuation and uncertainties in the results, it seems this
level of accuracy gives a good insight into decision making on well
4. Conclusion completion type, sand control selection, and field management.

In the present study, experimental and numerical literatures were Credit author statement
first reviewed and a wide range of laboratory conditions and many un­
certainties in numerical models were explained and categorized. The Mohammad Hossein Shahsavari: Conceptualization, Methodol­
main aim of this study was to compare the experimental data and nu­ ogy, Software, Writing- Original draft preparation. Ehsan Khamehchi:
merical sand production prediction model’s deficiencies and to show Supervision, Data curation, Writing- Original draft preparation. Vahi­
how simple parameters such as mesh size and its shape affect the results. doddin Fattahpour: Visualization, Investigation, Validation. Hamed
For this purpose, two widely used numerical models with Mohr- Molladavoodi: Software, Validation, Writing- Reviewing and Editing.
Coulomb elastic-perfectly plastic and Mohr-Coulomb cohesion soft­
ening/friction hardening behavioral models were applied. Moreover, Declaration of competing interest
the fracture energy regularization method was implemented to resolve
the mesh dependency in terms of finite energy dissipation and the The authors declare that they have no known competing financial
released energy amount of different mesh sizes. In addition, a mesh size interests or personal relationships that could have appeared to influence
sensitivity analysis was done to show the effects of mesh size and its the work reported in this paper.
shape and pattern on the results. Also, cumulative probability distribu­
tion versus absolute relative error diagrams were used to compare the
accuracy of the models. In the end, a model with less error was chosen to
simulate a real sand production in an oil well with 50◦ inclination and 6

22
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

References Geilikman, M., Dusseault, M., Dullien, F., 1994. Sand production as a viscoplastic
granular flow. In: SPE Formation Damage Control Symposium. Society of Petroleum
Engineers.
Abbas, A.K., Baker, H.A., Flori, R.E., Al-hafadhi, H., Al-haideri, N., 2019. Practical
Ghalambor, A., Hayatdavoudi, A., Alcocer, C.F., Koliba, R.J., 1989. Predicting sand
approach for sand-production prediction during production. In: 53rd US Rock
production in US gulf coast gas wells producing free water. J. Petrol. Technol. 41
Mechanics/Geomechanics Symposium. American Rock Mechanics Association.
(12), 1,336–1,343.
Aborisade, O., 2010. Practical Approach to Effective Sand Prediction, Control and
Haimson, B., Lee, H., 2004. Borehole breakouts and compaction bands in two high-
Management.
porosity sandstones. Int. J. Rock Mech. Min. Sci. 41 (2), 287–301.
Acock, A., et al., 2004. Practical approaches to sand management. Oilfield Rev. 16 (1),
Hall Jr., C., Harrisberger, W., 1970. Stability of sand arches: a key to sand control.
10–27.
J. Petrol. Technol. 22 (7), 821–829.
Al-Awad, M.N., El-Sayed, A.-A.H., Desouky, S.E.-D.M., 1999. Factors affecting sand
Ikporo, B., Sylvester, O., 2015. Effect of sand invasion on oil well production: a case
production from unconsolidated sandstone Saudi oil and gas reservoir. J. King Saud
study of Garon field in the Niger Delta. Int. J. Eng. Sci. 4 (5), 64–72.
Univ.-Eng. Sci. 11 (1), 151–172.
Jafarpour, M., et al., 2012. Determination of mobilized strength properties of degrading
Alquwizani, S.A., 2013. Three-dimensional elasto-plastic modeling of wellbore and
sandstone. Soils Found. 52 (4), 658–667.
perforation stability in poorly consolidated sands.
James, G., Shuling, H., Yoon-ha, L., David, S., Kenny, Y., 2001. Prediction of oil
Alsiny, A., Vardoulakis, I., Drescher, A., 1992. Deformation localization in cavity
production with confidence intervals. In: SPE Reservoir Simulation Symposium.
inflation experiments on dry sand. Geotechnique 42 (3), 395–410.
Jing, L., Stephansson, O., 2007. Fundamentals of Discrete Element Methods for Rock
Bažant, Z.P., Lin, F.B., Lippmann, H., 1993. Fracture energy release and size effect in
Engineering: Theory and Applications. Elsevier, p. 85.
borehole breakout. Int. J. Numer. Anal. Methods GeoMech. 17 (1), 1–14.
Khamehchi, E., Reisi, E., 2015. Sand production prediction using ratio of shear modulus
Beck, M., Rinaldi, A., Flemisch, B., Class, H., 2020. Accuracy of fully coupled and
to bulk compressibility (case study). Egypt. J. Petrol. 24 (2), 113–118.
sequential approaches for modeling hydro-and geomechanical processes. Comput.
Kim, A., Sharma, M., 2012. A predictive model for sand production in realistic downhole
Geosci. 24 (4), 1707–1723.
condition. In: 46th US Rock Mechanics/Geomechanics Symposium. American Rock
Behrmann, L., Willson, S., de Bree, P., Presles, C., 1997. Field implications from full-scale
Mechanics Association.
sand production experiments. In: SPE Annual Technical Conference and Exhibition.
Kooijman, A., Halleck, P., de Bree, P., Veeken, C., Kenter, C., 1992. Large-scale
Society of Petroleum Engineers.
laboratory sand production test. In: SPE Annual Technical Conference and
Belheine, N., Plassiard, J.-P., Donzé, F.-V., Darve, F., Seridi, A., 2009. Numerical
Exhibition. Society of Petroleum Engineers.
simulation of drained triaxial test using 3D discrete element modeling. Comput.
Kulatilake, P., Malama, B., Wang, J., 2001. Physical and particle flow modeling of jointed
Geotech. 36 (1–2), 320–331.
rock block behavior under uniaxial loading. Int. J. Rock Mech. Min. Sci. 38 (5),
Bentley, R.W., 2002. Global oil & gas depletion: an overview. Energy Pol. 30 (3),
641–657.
189–205.
Lv, M., 2018. AN ANALYTICAL AND EXPERIMENTAL STUDY ON THE ONSET OF SAND
Bratli, R.K., Risnes, R., 1981. Stability and failure of sand arches. Soc. Petrol. Eng. J. 21
PRODUCTION: A PORO-ELASTOPLASTIC MODEL WITH STRAIN SOFTENING
(2), 236–248.
BEHAVIOUR.
Burton, R., Davis, E., Morita, N., 1998. Application of reservoir strength characterization
Mahmud, H.B., Leong, V.H., Lestariono, Y., 2020. Sand production: a smart control
and formation failure modeling to analyze sand production potential and formulate
framework for risk mitigation. Petroleum 6 (1), 1–13.
sand control strategies for a series of North Sea gas reservoirs. In: SPE Annual
Meier, T., Rybacki, E., Reinicke, A., Dresen, G., 2013. Influence of borehole diameter on
Technical Conference and Exhibition. Society of Petroleum Engineers.
the formation of borehole breakouts in black shale. Int. J. Rock Mech. Min. Sci. 62,
Carter, J.N., Ballester, P.J., Tavassoli, Z., King, P.R., 2006. Our calibrated model has poor
74–85.
predictive value: an example from the petroleum industry. Reliab. Eng. Syst. Saf. 91
Morita, N., 1994. Field and laboratory verification of sand-production prediction models.
(10–11), 1373–1381.
SPE Drill. Complet. 9 (4), 227–235.
Chin, L., Ramos, G., 2002. Predicting volumetric sand production in weak reservoirs. In:
Morita, N., Whitfill, D., Massie, I., Knudsen, T., 1989. Realistic sand-production
SPE/ISRM Rock Mechanics Conference. Society of Petroleum Engineers.
prediction: numerical approach. SPE Prod. Eng. 4 (1), 15–24.
Choi, J.-W., 2011. Geomechanics of Subsurface Sand Production and Gas Storage.
Mühlhaus, H., Vardoulakis, I., 1987. The thickness of shear bands in granular materials.
Georgia Institute of Technology.
Geotechnique 37 (3), 271–283.
Christie, M., Subbey, S., Sambridge, M., 2002. Prediction under uncertainty in reservoir
Naderi, M., Khamehchi, E., Karimi, B., 2019. Novel statistical forecasting models for
modeling. In: ECMOR VIII-8th European Conference on the Mathematics of Oil
crude oil price, gas price, and interest rate based on meta-heuristic bat algorithm.
Recovery.
J. Petrol. Sci. Eng. 172, 13–22.
Climent Pera, N., 2016. A Coupled CFD-DEM Model for Sand Production in Oil Wells.
Nouri, A., Kuru, E., Vaziri, H., 2009. Elastoplastic modelling of sand production using
Coates, G.R., Denoo, S., 1981. Mechanical properties program using borehole analysis
fracture energy regularization method. J. Can. Petrol. Technol. 48 (4), 64–71.
and Mohr’s circle. In: SPWLA 22nd Annual Logging Symposium. Society of
Nouri, A., Vaziri, H., Belhaj, H., Islam, R., 2004. Sand production prediction: a new set of
Petrophysicists and Well-Log Analysts.
criteria for modeling based on large-scale transient experiments and numerical
Cook, J., Bradford, I., Plumb, R., 1994. A study of the physical mechanisms of sanding
investigation. In: SPE Annual Technical Conference and Exhibition. Society of
and application to sand production prediction. In: European Petroleum Conference.
Petroleum Engineers.
Society of Petroleum Engineers.
Nouri, A., Vaziri, H., Kuru, E., Islam, R., 2006. A comparison of two sanding criteria in
Crook, T., Willson, S., Yu, J.G., Owen, R., 2003. Computational modelling of the
physical and numerical modeling of sand production. J. Petrol. Sci. Eng. 50 (1),
localized deformation associated with borehole breakout in quasi-brittle materials.
55–70.
J. Petrol. Sci. Eng. 38 (3–4), 177–186.
Nouri, A., Vaziri, H.H., Belhaj, H.A., Islam, M.R., 2007. Comprehensive transient
Cundall, P., Board, M., 1988. A microcomputer program for modelling large-strain
modeling of sand production in horizontal wellbores. SPE J. 12 (4), 468–474.
plasticity problems. In: PROCEEDINGS OF THE SIXTH INTERNATIONAL
Ogunkunle, F.T., Isehunwa, S., Orodu, O., Ifeanyi, S., 2018. Uncertainty assessment of
CONFERENCE ON NUMERICAL METHODS IN GEOMECHANICS, 11-15 APRIL 1988,
onset sand prediction model for reservoir applications. Cog. Eng. 5 (1), 1499580.
INNSBRUCK, AUSTRIA, vols. 1–3. Publication of: Balkema (AA).
Osisanya, S.O., 2010. Practical guidelines for predicting sand production. In: Nigeria
Das, B.M., 2013. Advanced Soil Mechanics. Crc Press.
Annual International Conference and Exhibition. Society of Petroleum Engineers.
Dean, R.H., Gai, X., Stone, C.M., Minkoff, S.E., 2006. A comparison of techniques for
Pamin, J., 1994. Gradient-dependent Plasticity in Numerical Simulation of Localization
coupling porous flow and geomechanics. SPE J. 11 (1), 132–140.
Phenomena (Ph. D. Thesis).
Deng, F., et al., 2019. Influence of sand production in an unconsolidated sandstone
Papamichos, E., Malmanger, E.M., 1999. A sand erosion model for volumetric sand
reservoir in a deepwater gas field. J. Energy Resour. Technol. 141 (9).
predictions in a North Sea reservoir. In: Latin American and Caribbean Petroleum
Desrues, J., Hammad, W., 1989. Shear banding dependency on mean stress level in sand.
Engineering Conference. Society of Petroleum Engineers.
In: Proc. Of the Int. Workshop on Numerical Methods for Localization and
Papamichos, E., Vardoulakis, I., Tronvoll, J., Skjaerstein, A., 2001. Volumetric sand
Bifurcation of Granular Bodies, pp. 57–67.
production model and experiment. Int. J. Numer. Anal. Methods GeoMech. 25 (8),
Eringen, A.C., 1966. A unified theory of thermomechanical materials. Int. J. Eng. Sci. 4
789–808.
(2), 179–202.
Papanastasiou, P.C., Vardoulakis, I.G., 1992. Numerical treatment of progressive
Ewy, R., Ray, P., Bovberg, C., Norman, P., Goodman, H., 1999. Openhole stability and
localization in relation to borehole stability. Int. J. Numer. Anal. Methods GeoMech.
sanding predictions by 3D extrapolation from hole collapse tests. In: SPE Annual
16 (6), 389–424.
Technical Conference and Exhibition. Society of Petroleum Engineers.
Pietruszczak, S., Mroz, Z., 1981. Finite element analysis of deformation of strain-
Fattahpour, V., Moosavi, M., Mehranpour, M., 2012. An experimental investigation on
softening materials. Int. J. Numer. Methods Eng. 17 (3), 327–334.
the effect of rock strength and perforation size on sand production. J. Petrol. Sci.
Pyatigorets, A.V., Russell, D.B., 2020. Implementation of Advanced Numerical Solvers in
Eng. 86, 172–189.
FLAC3D Thermal and Fluid Implicit Formulation.
Fuh, G.-F., Nozaki, M., 2014. Completion design using sand management approach based
Qiu, K., et al., 2006. Practical approach to achieve accuracy in sanding prediction. In:
on sanding prediction analysis for HPHT gas wells. In: SPE Annual Technical
SPE Asia Pacific Oil & Gas Conference and Exhibition. Society of Petroleum
Conference and Exhibition. Society of Petroleum Engineers.
Engineers.
Gazzola, L., et al., 2019. Uncertainty quantification and reduction through Data
Rahmati, H., et al., 2013. Review of sand production prediction models. J. Petrol. Eng.
Assimilation approaches for the geomechanical modeling of hydrocarbon reservoirs.
2013.
In: 53rd US Rock Mechanics/Geomechanics Symposium. American Rock Mechanics
Rahmati, H., Nouri, A., Vaziri, H., Chan, D., 2012. Validation of predicted cumulative
Association.
sand and sand rate against physical-model test. J. Can. Petrol. Technol. 51 (5),
Geertsma, J., 1985. Some rock-mechanical aspects of oil and gas well completions. Soc.
403–410.
Petrol. Eng. J. 25 (6), 848–856.

23
M.H. Shahsavari et al. Journal of Petroleum Science and Engineering 207 (2021) 109147

Ranjith, P., Perera, M., Perera, W., Choi, S., Yasar, E., 2014. Sand production during the Veeken, C., Davies, D., Kenter, C., Kooijman, A., 1991. Sand production prediction
extrusion of hydrocarbons from geological formations: a review. J. Petrol. Sci. Eng. review: developing an integrated approach. In: SPE Annual Technical Conference
124, 72–82. and Exhibition. Society of Petroleum Engineers.
Ranjith, P., Perera, M., Perera, W., Wu, B., Choi, S., 2013. Effective parameters for sand Vermeer, P.A., De Borst, R., 1984. Non-associated plasticity for soils, concrete and rock.
production in unconsolidated formations: an experimental study. J. Petrol. Sci. Eng. Heron 29 (3), 1984.
105, 34–42. Vriezen, P., Spijker, A., Van der Vlis, A., 1975. Erosion of perforation tunnels in gas wells.
Risnes, R., Bratli, R.K., Horsrud, P., 1982. Sand stresses around a wellbore. Soc. Petrol. In: Fall Meeting of the Society of Petroleum Engineers of AIME. Society of Petroleum
Eng. J. 22 (6), 883–898. Engineers.
Rösler, J., Harders, H., Bäker, M., 2007. Mechanical Behaviour of Engineering Materials: Walton, I.C., Atwood, D.C., Halleck, P.M., Bianco, L.C., 2001. Perforating unconsolidated
Metals, Ceramics, Polymers, and Composites. Springer Science & Business Media. sands: an experimental and theoretical investigation. In: SPE Annual Technical
Sambridge, M., Mosegaard, K., 2002. Monte Carlo methods in geophysical inverse Conference and Exhibition. Society of Petroleum Engineers.
problems. Rev. Geophys. 40 (3), 3-1-3-29. Wang, H., Sharma, M.M., 2018. The role of elasto-plasticity in cavity shape and sand
Stein, N., Hilchie, D., 1972. Estimating the maximum production rate possible from production in oil and gas wells. SPE J. 24 (02). Paper Number: SPE-187225-PA.
friable sandstones without using sand control. J. Petrol. Technol. 24 (9), 1, 157- Wehunt, C.D., 2003. Well performance with operating limits under reservoir and
1,160. completion uncertainties. In: SPE Annual Technical Conference and Exhibition.
Subbey, S., Christie, M., Sambridge, M., 2004. Prediction under uncertainty in reservoir Society of Petroleum Engineers.
modeling. J. Petrol. Sci. Eng. 44 (1–2), 143–153. Weingarten, J., Perkins, T., 1995. Prediction of sand production in gas wells: methods
Subbiah, S.K., et al., 2020. Root cause of sand production and methodologies for and Gulf of Mexico case studies. J. Petrol. Technol. 47 (7), 596–600.
prediction. Petroleum. https://doi.org/10.1016/j.petlm.2020.09.007. White, J.A., Castelletto, N., Tchelepi, H.A., 2016. Block-partitioned solvers for coupled
Taron, J., Elsworth, D., Min, K.-B., 2009. Numerical simulation of thermal-hydrologic- poromechanics: a unified framework. Comput. Methods Appl. Mech. Eng. 303,
mechanical-chemical processes in deformable, fractured porous media. Int. J. Rock 55–74.
Mech. Min. Sci. 46 (5), 842–854. Wu, B., et al., 2010. Sand production prediction for a mature oil field-a case study. In:
Terzaghi, K., 1936. Stress Distribution in Dry and in Saturated Sand above a Yielding SPE Asia Pacific Oil and Gas Conference and Exhibition. Society of Petroleum
Trap-Door. Engineers.
Tixier, M., Loveless, G., Anderson, R., 1975. Estimation of formation strength from the Wu, B., Choi, S., 2012. Effect of mechanical and physical properties of rocks on post-
mechanical-properties log (incudes associated paper 6400). J. Petrol. Technol. 27 failure cavity development-experimental and numerical studies. In: 46th US Rock
(3), 283–293. Mechanics/Geomechanics Symposium. American Rock Mechanics Association.
Trampert, J., 1998. Global seismic tomography: the inverse problem and beyond. Inverse Wu, B., et al., 2016. A new and practical model for amount and rate of sand production
Probl. 14 (3), 371. estimation. In: Offshore Technology Conference Asia. Offshore Technology
Tronvoll, J., Skj, A., Papamichos, E., 1997. Sand production: mechanical failure or Conference.
hydrodynamic erosion? Int. J. Rock Mech. Min. Sci. 34 (3–4), 291 e1-291. e17. Wu, B., Tan, C.P., Lu, N., 2005. Effect of water-cut on sand production-an experimental
Unander, T., Papamichos, E., Tronvoll, J., Skj, A., 1997. Flow geometry effects on sand study. In: SPE Asia Pacific Oil and Gas Conference and Exhibition. Society of
production from an oil producing perforation cavity. Int. J. Rock Mech. Min. Sci. 34 Petroleum Engineers.
(3–4), 293 e1-293. e15. Yoshida, T., 1995. Shear banding in sands observed in plane strain compression. Localiz.
Van den Hoek, P., Geilikman, M., 2003. Prediction of sand production rate in oil and gas Bifurcation Theor. Soils and Rock 165–179.
reservoirs. In: SPE Annual Technical Conference and Exhibition. Society of Zeng, Y., Zhang, B., Wang, Y., 2016. Thermal effects on sand prediction. In: 50th US Rock
Petroleum Engineers. Mechanics/Geomechanics Symposium. American Rock Mechanics Association.
Van den Hoek, P., et al., 1996. A new concept of sand production prediction: theory and Zhu, H., et al., 2020. Numerical analysis of sand production during natural gas extraction
laboratory experiments. In: SPE Annual Technical Conference and Exhibition. from unconsolidated hydrate-bearing sediments. J. Nat. Gas Sci. Eng. 103229.
Society of Petroleum Engineers. Zivar, D., Shad, S., Foroozesh, J., Salmanpour, S., 2019. Experimental study of sand
Van den Hoek, P., et al., 2000. Horizontal-wellbore stability and sand production in production and permeability enhancement of unconsolidated rocks under different
weakly consolidated sandstones. SPE Drill. Complet. 15 (4), 274–283. stress conditions. J. Petrol. Sci. Eng. 181, 106238.
van den Hoek, P.J., Geilikman, M.B., 2005. Prediction of sand production rate in oil and
gas reservoirs: field validation and practical use. In: SPE Annual Technical
Conference and Exhibition. Society of Petroleum Engineers.

24

You might also like