Download as pdf or txt
Download as pdf or txt
You are on page 1of 37

Journal Pre-proof

Criterion of proppant pack mobilization by filtrating fluids: Theory and experiments

Dimitry Chuprakov, Aliia Iuldasheva, Alexey Alekseev

PII: S0920-4105(20)30853-6
DOI: https://doi.org/10.1016/j.petrol.2020.107792
Reference: PETROL 107792

To appear in: Journal of Petroleum Science and Engineering

Received Date: 10 January 2020


Revised Date: 19 May 2020
Accepted Date: 12 August 2020

Please cite this article as: Chuprakov, D., Iuldasheva, A., Alekseev, A., Criterion of proppant pack
mobilization by filtrating fluids: Theory and experiments, Journal of Petroleum Science and Engineering
(2020), doi: https://doi.org/10.1016/j.petrol.2020.107792.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier B.V.


CRediT author statement
Dimitry Chuprakov: Conceptualization, Methodology, Investigation, Writing- Original draft
preparation, Writing- Reviewing and Editing, Formal Analysis, Visualization

Aliia Iiuldasheva: Methodology, Software, Validation, Investigation, Formal Analysis,


Writing- Original draft preparation, Writing- Reviewing and Editing, Visualization

Alexey Alekseev: Conceptualization, Methodology, Visualization, Investigation, Validation.

of
ro
-p
re
lP
na
ur
Jo
Criterion of proppant pack mobilization by
filtrating fluids: theory and experiments
Dimitry Chuprakov1, Aliia Iuldasheva2
Schlumberger Moscow Research, 13 Pudovkina St., Moscow, Russia, 119285
Alexey Alekseev3
Schlumberger Novosibirsk Technology Center, 1/10 Zelyonaya Gorka,
Novosibirsk, Russia, 630060

of
ABSTRACT

ro
The goal of hydraulic fracturing is placing a granular material, proppant, in the created
-p
fracture to sustain its conductivity and attain economical production rates. After fracture
closure on the placed proppant and subsequent pressure drawdown in a well, the consolidated
re
proppant can be dragged by the producing fluids and undesirably be produced back to the well.
Rough estimations or accurate simulations of this phenomenon require a valid quantitative
lP

criterion of proppant mobilization in a fracture.


In this work, we analytically derive such criterion from the fundamental principles. The
criterion is expressed in terms of the minimum fluid filtration velocity that destabilizes the
na

proppant. It is sensitive to such important factors as cohesion, thickness and permeability of a


proppant pack, mean grain size, roughness and softness of fracture walls, and filtrating fluid
ur

viscosity. The developed theory considers the existing unloaded edge of a proppant pack, its
granular structure, and related finite cohesion of particles holding each other in a pack. These
Jo

key properties are shown to play a dominant role for mobilization of a proppant in hydraulic
fractures.
On the other hand, we conduct an extensive series of laboratory experiments mimicking
subterranean fracture closure on proppant, fluid filtration through it, and registering proppant
flowback. In the experiments, we try a wide range of commercial proppants, less- and more-
viscous fluids, and stiff and soft fracture walls. We associated proppant flowback onset with the
change of proppant pack height. We also show how proppant bridging depends on the particle
embedment into soft walls and illustrate various geometries of proppant washing out, giving
clues to various scenarios of proppant flowback evolution.
Finally, we observe a good correspondence of our theory and experimental observations.
Theoretical hypotheses of unstressed edge and finite cohesion of a granular solids serve as the
only means to explain the observed range of filtration velocity magnitudes for proppant

1
Corresponding author, email: dchuprakov@slb.com
2
Email: agazizova@slb.com
3 Email: AAlekseev4@slb.com

1|Page
mobilization and their independence from the value of confining stress. The reciprocity between
the cohesion and critical velocity revealed by the theory is confirmed by experiments. It allows
measuring only one of them to estimate another.

KEYWORDS:
Proppant flowback; Dense granular suspension; Sand mobilization; Hydraulic fracture;
Bridging.

1. INTRODUCTION
Proppant flowback or solid production is not a novel topic in oilfield research and

of
engineering. It has been addressed and studied for several decades (Asgian et al., 1995; Milton-
Tayler et al., 1992). The literature includes publications describing field observations (Elliott

ro
and Gale, 2018; Potapenko et al., 2019, 2017), laboratory experiments (Guo and Ugwu, 2015;
McLennan et al., 2015; Zhang and Guo, 2016) and numerical (Asgian et al., 1995; Parker et al.,
-p
1999) and analytical models (Canon et al., 2003; Cerasi and Mills, 1998; Pearson and Zazovsky,
1997) aiming at describing proppant mobilization inside a fracture under the applied normal
re
stress. A realistic fracturing treatment is subject to nonuniform proppant placement across the
fracture area (Morris et al., 2014; Velikanov et al., 2018). Some zones of the fracture area
lP

accumulate thicker proppant packs whereas other zones can have either poorer proppant
concentration or contain no proppant at all (Velikanov et al., 2018). As a result, the created
na

hydraulic fracture can be closed on proppant heterogeneously by heavily pressing thick


proppant pack areas and leaving poorly propped parts of a fracture unclamped (Morris et al.,
2014; Pyrak-Nolte and Morris, 2000). After injection into fractures, proppant may occur in two
ur

different mechanical states: a dilute suspension and a dense pack of solid particles. The first
state is similar to a slurry being injected into a fracture during the hydraulic fracturing
Jo

treatment. The second state represents a thin layer of solid porous medium in which the
proppant is consolidated under the confining stress and yet is weakly cohesive and easily
damageable. Both states of proppant can be described by one general theory because they obey
the same equations of mechanics and refer to one type of suspension. However, following
certain application intentions, previous studies often considered only one of those states as a
principal one and neglected proper analysis of the other.

A fundamental theory of monodisperse granular suspension flow subjected to a confining


stress has been developed (Guazzelli and Pouliquen, 2018). It is based on the fundamental laws
of mass and momentum conservations and complemented by a few material closure relations
obtained from experiments. In one such experimental study (Boyer et al., 2011), the authors
found unified rheology laws coupling the shear dilation (shear thickening) of the suspension
with the applied shear and normal stresses. As a result of this work, one can associate the
volumetric concentration of the granular solids with the applied normal and shear stresses. On
the other hand, the rheology of suspension flow is expressed in terms of the shear rate as a
function of solid particle concentration and applied stresses. Both relationships taken together
help to build a fully coupled model of proppant-laden fluid flow in a fracture (Dontsov and
Peirce, 2014; Lecampion and Garagash, 2014). The results of the suspension theory above were

2|Page
not intentionally designed for the oilfield industry or for hydraulic fractures. However, they
were used in construction of the advanced models for proppant placement during hydraulic
fracturing treatments. Using those results, Dontsov and Peirce (2014) obtained a solution for
the Poiseuille problem of a steady-state suspension flow in a vertical fracture channel. Dontsov
and Peirce obtained a solution for velocities of fluid and solid components in a fracture given
the pressure gradient as a function of solid concentration. In this solution, the authors clearly
showed a transition between the classical Poiseuille flow and the Darcy flow regime as proppant
concentration went from small to large values. Lecampion and Garagash (2014) revisited the
compressibility law for a suspension reported by Boyer et al. (2011). Based on the results of
additional experimental data, Lecampion and Garagash (2014) considered further compaction
of the suspension beyond the jamming limit. Nevertheless, the models by Dontsov and Peirce
(2014) and Lecampion and Garagash (2014) look similar. They use many identical assumptions,
and both primarily target the scenario of dilute slurry injection into an open channel. Their
models might need to be enhanced to include the possibility of mobilization of proppant already

of
consolidated into packs.

ro
Another class of studies refers to understanding of the proppant flowback specifically.
These studies pay special attention to the mobilization of firmly compacted solids and ignore
-p
poorly confined suspension cases considered by the studies mentioned above. Though, as
mentioned, both highly and weakly compressed solids may simultaneously exist inside a
re
heterogeneously propped fracture. First, there were multiple laboratory experiments that noted
key aspects of proppant destabilization during flowback and production. Milton-Tayler et al.
lP

(1992) identified key parameters that cause proppant flowback: fracture width, applied
pressure gradient and confining stress. They also reported limited influence of the confining
stress on proppant flowback. Asgian et al. (1995) investigated key motive factors for proppant
na

pack mobilization and instability by means of the explicit numerical modeling of proppant
particles by the distinct element method (DEM). In DEM modeling, Asgian handled several
ur

thousands of particles co-interacting by normal and shear forces at each particle-particle


contact. This work concluded that a cohesionless, unbonded proppant pack under closure stress
Jo

inside a hydraulic fracture becomes unstable at a certain critical ratio of fracture width to mean
proppant grain size. It also numerically determined that the critical drawdown pressure
gradient to cause the proppant pack destabilization is 75 psi/ft. Erosional washing out of
proppant pack from edges has been experimentally observed and modelled by Parker et al.
(1999). They showed the formation of proppant-free finger-like channels, the geometry of
which varies depending on the fluid flowback rate.

The edge of a proppant pack deserves special attention. This is the zone where the normal
stress must be weakened due to the presence of a free unsupported boundary at the edge.
However, the edge of consolidated granular solids can still sustain some level of stress that
holds the proppant pack if the latter exhibits finite cohesion. It was experimentally shown that
wet sands exhibit a nonzero cohesion (Richefeu et al., 2006). The intrinsic cohesion of granular
materials is shown to be possible due to their granular structure, as well as existing capillary
forces in the water wetting each particle (Than et al., 2017). In addition to multiple empirical
relations quantifying the bridging effect, Garagash et al. (2019) developed a mathematical
theory of spherical particle bridging in a fracture.

Extensive numerical studies enabled better understanding of the phenomenon of


proppant flowback, and fast analytical models facilitated deployment of these models in

3|Page
practice. In this work, we develop a simple and useful analytical model that can be used in
engineering and commercial applications. In our model, we consider most of the noticeably
important effects reported in earlier publications: thickness of a pack and its cohesion, size of
particles and bridging effect, role of the proppant pack edge, wall roughness and particle
embedment into walls and others. The model features a good balance between the simplicity of
implementation and sensitivity to various important parameters. That gives it a chance to be
selected for integration into the commercial simulators of fluid flow in propped fractures.

The structure of this work is the following. First, we formulate the mathematical problem
and obtain the solution for stresses in three distinctive regions near a proppant pack edge.
Using this solution, we derive a critical filtration velocity that initiates proppant movement at
the edge of a pack. We also propose how the theory can quantify the bridging factor by the
fracture wall roughness and softness. Next, we describe the laboratory experiments and
observations of different proppant flowback scenarios. The final section is devoted to validation

of
of the theoretical criterion for proppant mobilization by the results of laboratory experiments.
We discuss reciprocity of the critical filtration velocity and cohesion of the washing out granular

ro
structures. A summary of the work achievements is given in the Conclusions section of the
paper.
-p
re
2. THEORY
lP

2.1. The problem


We consider the semi-infinite layer of a granular proppant pack placed between two parallel flat
na

horizontal plates, as shown in Fig. 1. The infinite and rigid (nondeforming) plates squeeze a
compressible proppant pack by the prescribed confining stress and ensure that the width
ur

is uniform everywhere along the pack. Additional to loading, an incompressible Newtonian


fluid with viscosity is filtrating with velocity from the far-field region (Fig. 1, right-hand
Jo

side) towards the edge (Fig. 1, left-hand side).

4|Page
Figure 1. Semi-infinite proppant pack with free edge at = 0 subjected to the external normal load and
fluid filtration along OX with velocity (right to left). The edge results in nonuniform proppant
pack loading at its proximity with the weakest solid particle pressure at the edge, followed by
the monotonically increasing solid particle pressure in the near-edge region, and constant
far-field value . The dashed lines on the right denote infinite extension of the compressed pack
to the right-hand side. The red arrows in the mobile region indicate possibility of solid
mobilization from the edge.

Choose OX axis such that it propagates along the pack and the plates, and its origin, = 0,

and = 0 coincides with the middle of a pack layer (Fig. 1). Filtration velocity
corresponds to the position of a proppant pack edge. The OY axis is directed across the pack,
and velocity
of a mobilized solid are co-directed with the OX axis and have negative sign. The pressure of
solid proppant particles in the pack is not uniform due presence of the edge. It is weakest
at the edge , followed by its monotonic increase in the near-edge region , and
maximum value at the far-field region.

of
The proppant is considered as a porous, permeable and perfectly plastic material, which has a

ro
finite cohesion (i.e., finite yield stress) but is loaded beyond the yield stress (Fjaer et al.,
2008; Jaeger et al., 2007). If mobilized, the solid particles will move from the edge to the left-
hand side (mobile region) with velocity .
-p
re
2.2. Stress state in solid-fluid mixture
lP

The mobility of solids and fluids is determined by their stress state. The partial stress
tensor for solid particles
na


= (1)
ur

must be added up with the partial fluid stress tensor


Jo

= (2)

to get the total stress applied to the suspension of solid particles:

+ − +
= = + = + + (3)

Here, and are the positive fluid pressure and transverse normal stress applied to a particle
pack, respectively. Similarly, and , denote shear stresses applied to the fluid and solid
components, respectively. For solid particles, modelled as a perfectly plastic material, we
employ the Tresca criterion (Fjaer et al., 2008; Jaeger et al., 2007), which defines the following
relationship between the longitudinal ( ) and transverse ( ) normal stresses in the pack:

= − (4)

5|Page
where is the uniaxial yield stress. It has nonzero value for granular frictional materials and
has the same order of magnitude as the cohesion (Fjaer et al., 2008; Jaeger et al., 2007).
In this work, we use the following convention for the sign of stress components. Normal stress
components , are positive (Fig. 2, left), when compressive, and, vice versa, they are
negative if they produce tension. Similarly, the shear stress is taken positive if it yields
rotation in a counterclockwise direction (Fig. 2, right) and negative when the rotation is in
clockwise direction.

of
ro
-p
re
Figure 2. Positive sign for normal (left) and shear (right) stresses.
lP

2.3. Proppant pack mobilization


na

The proppant pack can be in two steady states: immobile or mobile. In the immobile state,
the proppant particles do not move even when the fluid is filtrating through them. In the mobile
state, the proppant pack moves when subjected to shear stress like a fluid. But, unlike a fluid,
ur

the shear flow of a granular suspension is initiated only when its local concentration is below
the critical concentration corresponding to the onset of jamming, i.e.
Jo

(5)

It is experimentally shown that for a granular suspension subjected to fixed normal stress, the
shear stress proportionally reduces concentration of the suspension (Boyer et al., 2011;
Guazzelli and Pouliquen, 2018). The rheology measurements allow rewriting the criterion (5) in
terms of the critical shear stress that mobilizes suspension (Boyer et al., 2011):

! (6)

where =" with the coefficient of friction " = 0.32, as experimentally found by Boyer
(2011). The shear stress can be found by solving the following momentum equations, written
in the absence of gravitation force, as

& =& '


+& (
=) (7)

which, after integration over across a channel, give

6|Page
= + = −*∇ +∇ , = −∇ ∙ (8)

Here we assume that . and . are constant across the proppant pack, and thus shear stress
is a linear function of . This assumption is justified because the fracture channel is much

Hereafter for brevity, we denote . = / ⁄/ , ∇ = / ⁄/ and ∇ = ∇ + ∇ is the total


thinner than any typical scale of fluid and solid pressure changes along the fracture channel.

pressure drop. As there is no evidence how significant the flow of dense suspension inside a thin
channel with or without wall roughness will deviate from laminar flow, we follow the

shear stress is reached at the wall ( = /2). Then, Boyer’s criterion of proppant pack
assumption of laminar flow inside a channel with flat parallel walls, where the maximum of

mobilization (6) reads

= −∇ ∙ !"
2
2

of
(9)

ro
As far as this criterion has been developed for a large number of particles per thickness of a
pack, in this work, we introduce correction to this criterion that considers the granular
-p
structure of proppant packs. We note that to associate the shear stress applied to a pack of solid
particles with their concentration , the stress must be averaged over some representative
re
averaging must be performed over some 3/ scale, where 3 ! 1 but 5 1 , assuming in what
volume of a suspension. The volume must include, at least, several particles. Hence, the stress

follows that > 3/. Particularly, the maximum shear stress 2 at the wall must be replaced by
lP

the averaged shear stress 2,〈 9〉 , obtained using (8), as


2/=
1 3/
na

= ; ′ / ′ = *& +& , 1−
2,〈 9〉
3/ 2 (10)
2
=> 9
ur

correction factor 1 − 3// with respect to the local shear stress 2 at the wall:
This result shows that averaging over a small ensemble of particles results in the additional
Jo

3/
2,〈 9〉 = 2 ?1 − @ (11)

mean proppant size /, then 3/⁄ ≪ 1 and 2,〈 9〉 = 2 , as intrinsically assumed in multiple
This result reduces to the criterion by Boyer; if the fracture width is much larger than the

suspension studies (Guazzelli and Pouliquen, 2018).


So, replacing the shear stress (10) in (9), the corrected criterion of proppant mobilization in
terms of pressure and stresses is written as

3/
= *& +& , 1− !"
2,〈 9〉
2
(12)

where we note that 2,〈 9〉 is the absolute value of shear stress and is a non-negative value.

7|Page
2.4. Fluid filtration through solids
It is practically more convenient to express the proppant mobilization criterion (12) in
terms of velocities of fluid filtration rather than shear stresses or pressure gradients. We
assume that the velocity of fluid filtration obeys the Darcy law

B
= ∇ (13)

where B is the permeability of a proppant pack, and it is assumed that has a positive sign in
the considered problem:

C = *− , 0, 0, (14)

of
as the fluid flows from the right-hand side to the left-hand side (Fig. 1), and thus, according to
the chosen direction of OX axis, is negative.

ro
The criterion of mobilization (12) can be written in terms of the minimum filtration velocity
that causes proppant to move:

B 2"
-p
= ? −& @
,D
− 3/
re
(15)
lP

The stresses applied to a proppant pack are not uniform along the longitudinal axis near
the edge. We recognize the three regions: the far-field, near-edge and edge, which differ in the
stress state and hence in ability to mobilize the proppants. For each region, below we evaluate
na

the magnitude of filtration velocity (15), which is sufficient to mobilize it.


ur

2.5. Far-field region


Jo

The far-field region, which is sufficiently far away from the edge of a pack, is uniformly squeezed
by the confining walls (Fig. 1, right-hand side). Because of substantial size along OX, it fully
supports the load externally applied to the walls, and so

= (16)

where = − is constant. Assuming relatively large particle pressure, i.e., > ,


the longitudinal normal stress component is also constant:

= − (17)

where the Tresca criterion (4) is used. From the solution of momentum equation (8), it means
that there is no shear stress induced in the far-field region:

=0 (18)

8|Page
far-field region. As a practical example, consider a proppant pack of thickness = 5 mm with
Let us estimate the critical filtration velocity for mobilization of a proppant pack in the

mean grain size / = 1 mm and permeability B = 1000 D. Assume that the in-situ rock stress
= 30 MPa, fluid pressure inside a fracture = 20 Mpa and fluid
= 1 cP = 10 Pa ∙ s. Then, acting particle pressure = 10 MPa, and criterion
applied to fracture walls
>N

(15) gives the minimum value of filtration velocity to mobilize that region ,D = 2.5 km/s. This
viscosity is

value is huge with respect to normal filtration velocities of producing fluids in hydraulic
fractures, which are on the order of only centimeters per second. That means that mobilization
of proppant packs compressed by typical rock stresses can never occur in the packs’ middles. It
can only evolve from their edges.

2.6. Near edge region

of
Criterion (12) gives a notice though, that there is one more mechanism to cause mobility;

pressure gradient ∇ . This mechanism of mobilization occurs in the near-edge region, which is
this mechanism is not associated with fluid flow but is linked to the existence of particle

ro
-p
close enough to the edge of a pack. Due to a strong contrast between the weak edge stresses and
large stresses at the far-field region, proppant particles in this intermediate region partially

force of particle pressure gradient ∇ , which is proportional to the applied external stress, as it
re
squeeze out towards the edge (i.e., to the left-hand side in Fig. 1). This motion is driven by the
lP

will be clear from the following mathematical derivations (Eqs. (22)-(23)). As a result, that
squeeze-out partially unloads this region, and stresses there are lower than in the far-field
region (Fig. 1, left-hand side).
na

and, assuming that ∇ = constant, we find


To find the solution for stresses in the near-edge region, we integrate the equality of (12)

2" − 3/
ur

= T exp + ∇
− 3/ 2"
(19)
Jo

where the unknown coefficient T is found from the stress condition at the edge ( = 0):

− 3/
T= − ∇
W
2"
(20)

The extent of the near-edge region XYW is then found from the stress condition at the far-field
region ( = XYW ):

XYW = (21)

and it is
− 3/
− 3/ − 2" ∇
XYW = ln [ \
2" − 3/
− 2" ∇
W
(22)

The typical value of a near-edge region extent XYW can be estimated for the example given
above. Taking the particle pressure and properties of a proppant pack from that example and

9|Page
= 100 Pa (choice of this value will be clear
W

below), from (22) we find XYW = 4.5 cm.


assuming that the stress at the edge region is

For proppant packs loaded by typical magnitudes of rock stresses, the perturbation
created for the near-edge region by the filtrating fluids is negligibly small. So, Eq. (22) can be
simplified to

− 3/
XYW = ln ^ _
2" W
(23)

2.7. Edge region

of
a pack. That conditions make the edge region different from the near-edge region. At = 0, the
At the very edge, the stresses are dictated solely by the boundary conditions at the edge of

ro
longitudinal effective normal stress and shear stress are zero:

=0 =0
-p (24)
re
=0 =0 (25)
lP

From the Tresca criterion, it follows that the transverse normal stress at the edge is equal to the
na

uniaxial yield stress :

=0 = =
W
(26)
ur

which means that the edge region is compressed by the finite cohesion of a proppant pack.
Jo

Combining (8) and (25), we find that

∇ =0 (27)

= . We estimate its size as


W

5
which gives a definition of the edge region as the region where
. The stress conditions (24)-(27) are supposed to be held along the entire edge region.
The critical filtration velocity that can destabilize this region is written from (15), (26)
and (27) as

2B "
=
,D
/ −3
/
(28)

This is the minimum velocity that can destabilize the pack, because the particle pressure is
the smallest in this region.

= 100 Pa, we
W
Taking the example of the proppant pack above and assuming again that
obtain the critical filtration velocity that will cause mobilization of the proppant pack edge

10 | P a g e
,D = 2.5 cm/s. This value is realistic and can be achieved in real fluid flow conditions in a
fracture.
In the absence of fluid filtration though the pack, the pack stays stationary. Shear stress induced
by fluid filtration may destabilize the pack at the edge, which is the weakest part and so is most

width to particle diameter // is less or close to the bridging factor 3, as predicted by Eq. (28).
prone to destabilization. However, the mobilization will not occur if the ratio of proppant pack

2.8. Bridging enhancement by wall roughness and


softness
Proppant particle bridging is caused by geometric constraints, and, the technically
possible increase of fluid filtration velocity cannot destabilize the bridged proppant pack

of
pack // is in the range of 2 to 3.
(Garagash et al., 2019). These geometrical constraints arise when the number of particles per

ro
On the other hand, the bridging depends on the roughness of fracture walls. The
-p
roughness is associated with additional geometrical tortuosity of fracture walls, which intake a
certain part of proppants adjacent to the fracture walls and block their shear flow in the pack
re
hold the particles captured by wall roughness (Fig. 3, `). The result is that the real width of a
even if the significant shear stress is applied (Fig. 3). This occurs because wall surface asperities

fracture aWbc , where the shear proppant flow may take place, is less than the ideal width of a
lP

proppant pack deWbc , which we previously assumed. Let us denote difference between those
widths 2f [m], assuming proppant pack symmetry. The value of 2f can be assessed as an
na

averaged part of proppant pack width absorbed by wall roughness.


ur
Jo

a) b)
Figure 3. The influence of fracture wall roughness f (a) and particle embedment g (b) on proppant flowback.

Furthermore, the bridging is enhanced by softness of the rock walls that compress the

particles into the walls proportionally to the stress applied (Fig. 3, h). The higher the closure
proppant (Ghanizadeh et al., 2016; Li et al., 2015). Soft rock facilitates embedment of proppant

of particles embedment g [m], similar to the effect of roughness.


stress, the larger the particle embedment. The embedment can also be quantified by the depth

11 | P a g e
It is important that effective reduction of the fracture width must be accounted for in
the cases in which fracture wall roughness and embedment are essential. This can be a typical
situation for hydraulic fractures in rock. So, we introduce the following correction for the real
fracture width aWbc as opposed to the ideal one :

aWbc = − 2f − 2g (29)

Substituting now the real width aWbc in place of the ideal width to the criterion (28), we get

2B "
= ∙
,D
/ −h
/
(30)

where h is the bridging factor defined as

of
2f 2g
h=3+ +
/ /

ro
(31)

-p
noticeable increase of bridging factor h in our laboratory experiments when we replace the steel
This more complete criterion of proppant mobilization (30)-(31) allows us to explain a
re
confining plates with soft and rough polymer plates (see the following section). Future

embedment depths, f and g, respectively, for the specific rocks. These parameters may become
applications of the model (30)-(31) will need to evaluate or measure the roughness and
lP

as large as 3 (which is of order of 2-3 as mentioned) and, thus, show enhanced proppant
na

bridging in fractures in rocks as opposed to the presented experiments with steel plates.
ur

3. LABORATORY EXPERIMENTS
Jo

3.1. Description of experiments


Materials. Several lightweight (LWP) and intermediate-strength (ISP) ceramic proppants
and one fracturing sand (FS) were selected for testing and modeling of basic flowback
properties of proppants. This choice was governed by the fact that these types of propping
agents are typically used in the field. For instance, sands or in some cases ceramic proppants
with the mesh size 20/40 are typically pumped in the “tail-in” stage of fracturing jobs in North
America. Ceramic proppants of 12/18 mesh size are often used in the tail-in stage in Russia.
General properties of the proppants are presented in Table 1.

Proppant Type LWP LWP ISP FS


Mesh Size 20/40 12/18 12/18 20/40
Average Diameter, mm 0.736 1.408 1.348 0.567
3
Absolute Density, g/cm 2.70 2.70 3.23 2.65
3
Bulk Density, g/cm 1.55 1.63 1.92 1.58
Roundness 0.80 0.90 0.89 0.91
Sphericity 0.90 0.90 0.87 0.82
Table 1. Properties of tested proppants.

12 | P a g e
Equipment. Proppant flowback testing was performed on a flowback apparatus that
consisted of (Fig. 4)
• Flowback cell (see detailed description below)
• Hydraulic press equipped with heating system
• COMMET YA65 water pump
• 30-L water bath
• Set of sensors (thermocouples, flowmeters, one pressure and one depth gauge)
• Data acquisition system and computer

of
ro
-p
re
Figure 4. Proppant flowback apparatus (left), schematic of flow loop (right) where P1 stands for
lP

pressure gauge.
Fig. 5 shows a schematic of the proppant flowback cell with a working area of 133.4 ×
na

133.4 mm that consisted of


• Cell body
ur

• Cell bottom
• Moving piston
Jo

• Top/bottom steel plates (133-mm square, 10-mm deep)


• Inlet and outlet funnels (not presented in Fig. 5)
Prior to the testing, the flowback cell was water sealed with silicone rubber and rubber O-
rings. To test flowback stability, a proppant pack is sandwiched between two plates. Closure
stress is applied normally to the proppant pack. It is kept constant at set point for the duration
of the test by an electronically controlled ISCO 260D syringe pump. Tap water or a mixture of
tap water and glycerol (1:1 by volume) are heated in a dedicated water bath up to 40°C and
pumped through the flowback cell by a COMMET YA65 water pump with a step rate from 0.5 to
20 L/min (with constant rate increment of ~0.8 L/min and ramp slope at acceleration of ~1.5
L/min). Pumping pressure and fluid temperature are measured before the inlet face of the
proppant pack that allows monitoring the test conditions and calculating differential pressure.
Proppant pack height is measured by a Mitutoyo digital depth gauge that measures the distance
between piston and cell bottom in the middle of the flowback cell (Fig. 5). It is assumed that the
piston and the cell bottom are parallel to each other in the course of the test duration and
therefore the proppant pack height is measured at its highest point.
Proppant flowback may occur from an outlet, and free proppant flows down the outlet
funnel to a fine mesh metal sieve where it is collected for weighing.

13 | P a g e
In our studies, we took two extreme cases of rock hardness: very soft and very hard rocks.
The hard rocks were imitated by steel plates that were mounted on flowback cell bottom and
piston (Fig. 5). The soft rock was mimicked by applying slip-resistant resilient tapes (sandpaper
applied on a polymer adhesive tape), to the steel plates that were mentioned above.

of
ro
-p
Figure 5. Scheme of proppant flowback cell (left), 0.6-mm layer of polymer adhesive tape on a steel slab
re
(right).
lP

Proppant pack preparation. Prior to testing, the cell outlet open face is closed with an
outlet plug. Proppant is wetted with a small amount (~30 mL) of a tap water. Then, the
proppant is loaded into the cell and thoroughly leveled, so that the height difference between
na

opposite corners does not exceed 0.2 mm.


Experimental procedure and conditions. After the proppant sample is loaded into the
ur

cell and leveled and the piston is assembled, the closure stress is ramped up to the target value
(Table 2). The pack is left for 20 minutes to stabilize. After that time, the outlet plug is removed,
Jo

and the pack width is measured at the open face. Then the preheated fluid is pumped through a
bypass for an additional 10 minutes with the rate of 6 L/min to heat the pipe system to a
required temperature. After 30 minutes of total time of stress exposure, the bypass is closed,
and fluid flow through the cell is initiated at ~0.5 L/min and increased with a constant rate
increment of ~ 0.8 L/min and ramp slope at acceleration of ~ 1.5 L/min. Fluid flow is
maintained at each level for 1.5 minutes if no proppant pack failure occurs. The test is continued
until the proppant pack fails or the maximum pumping rate of ~ 20 L/min is reached.
Experimental conditions of all proppant flowback tests given in this work are presented in
Table 2.

Parameter Range of values


Proppant LWP 20/40, LWP 12/18, ISP 12/18, FS 20/40
Proppant loading, g 40 – 380
Temperature, °C 40
Stress, bar (psi) 6.9 – 276 (100 – 4000)
Pumping rate, L/ min 0.5 – 20
Pumping fluid Water, Glycerol + Water (1:1 vol.)
Cell walls material Steel, Sandpaper
Table 2. Experimental conditions of proppant flowback tests.

14 | P a g e
3.2. Criteria of proppant pack failure
Criteria of proppant pack failure. At present, there are no clear and industry-accepted
criteria of proppant pack failure in flowback experiments. But based on the practical experience,
we developed the following criteria of proppant pack failure:
• Drop of proppant pack height
• Drop of differential pressure
• Massive appearance of proppant in outlet funnel and on the sieve
For well-consolidated proppant packs with 2.5 < //< 3, all three criteria are fulfilled
simultaneously, as the experimental logs in Fig. 6 demonstrate at the 16th minute. Using this

of
behavior in the experimental logs, the first drop of height was selected as a moment of the
proppant pack failure. In this case, the proppant pack failure is usually related to end-to-end

ro
channel formation (Fig. 7).

-p
re
lP
na
ur
Jo

Figure 6. Experimental logs of differential pressure, pumping rate and proppant pack height at the failure of
well consolidated proppant pack, arrow marks a moment of proppant pack failure.

15 | P a g e
of
Figure 7. End-to-end channel formation in the well consolidated proppant pack ISP 12/18. Arrow shows the

ro
fluid flow direction.

For unconsolidated proppant packs with // > 3, the picture of pack failure is different.
-p
In this case, there is no clear “breakthrough” moment. Instead, we observed a gradual decrease
re
of proppant pack height without a rapid drop of differential pressure (Fig. 8). Presumably, this
happens because the drag force created by a filtrating fluid pulls the proppant pack as a single
lP

mass. Therefore, the proppant pack migrates from the cell to the outlet funnel without breaking
its integrity. The less amount of proppant remaining in the cell, the lower is the height of the
proppant pack. Here again, the first drop in the height was selected as an indication of the
na

proppant pack failure.


ur
Jo

Figure 8. Experimental logs of differential pressure, pumping rate and proppant pack height at the failure of
unconsolidated proppant pack, arrow marks a moment of proppant pack failure.

In the experiments, a proppant pack is considered as stable if the maximum pumping rate
of 20 L/min was achieved and the pack did not fail, i.e., there was no drop of proppant pack
height and differential pressure, and no massive proppant return was detected.

16 | P a g e
3.3. Discussion of Experimental Results
A comprehensive literature overview of the proppant flowback problem revealed the
main factors affecting proppant pack stability (Asgian et al., 1995; Milton-Tayler et al., 1992;
Parker et al., 1999; Shor and Sharma, 2014). For the conventional proppants, these factors are
as follows:

• // ratio
• Embedment at stress
• Proppant type
• Fluid viscosity
Fluid flowrate is often included in the list of such factors, but in our testing, the flowrate
was considered as the “response”, and it is therefore excluded from the list. To cover all these
factors, extensive testing at various conditions has been performed.

of
i/j ratio. In Fig. 9, we present a dependence of the fluid linear velocity at proppant pack

ro
( // ratio). Colored circles and diamonds represent the different datasets obtained in the
failure (critical filtration velocity) on fracture aperture normalized to the proppant grain size
-p
experiments with steel plates for FS 20/40 and LWP 12/18 at various closure stresses.
re
lP
na
ur
Jo

Figure 9. Dependency of critical filtration velocity on proppant pack width in mean proppant diameters for
LWP 12/18 and FS 20/40 at various stress with steel plates and fluid viscosity of 0.65 cP.

These proppants demonstrate the same pattern of the flowback performance:

• The pack is extremely stable when // ratio is less than 2.5, which is related to

With further increase of // ratio, the pack showed a transition from stable to unstable
proppant bridging.

over a limited range of // ratios (from 2.5 to 3).


• At // ratio above 3, the pack loses integrity at constant critical filtration velocity of
about 1.5 cm/s.

17 | P a g e
In general, this tendency is in good correspondence with the facts reported in Milton-
Tayler et al. (1992). Those authors concluded that the pack turns from stable to unstable state

significantly higher // ratios. Apparently, that happened due to difference in material of the
over a narrow band of fracture apertures, but in their work the proppant pack stabilized at

top/bottom plates sandwiching the proppant during the flowback experiment. In the Milton-
Tayler et al. study, this material was plaster or sandstone, which are softer than steel we used.
As a result, proppant embedment in the plaster and sandstone is higher than in steel plates, and
that affects proppant pack stability.

Embedment at stress. The experimental fact that proppant pack strength is insensitive
to closure stress in the experiments with the steel plates is a bit controversial, but generally it
fits with the importance of embedment theory expressed in Milton-Tayler et al. (1992). Milton-
Tayler et al. stated that the stress affects the stability of the proppant pack, and it is primarily
related to embedment and rock hardness. Apparently, in the experiments with the steel plates,

of
the stress dependence was suppressed due to negligible embedment in the steel.

Another aspect of the insensitivity to closure stress that must be considered is the feature

ro
of the flowback experiment in which the fracture opening is situated horizontally. In this case,
the extraction of the outlet plug before starting the experiment leads to collapsing of the
-p
proppant pack edge (loose-end) and formation of the pack slope. As a result, the closure stress
drops to zero in this part of the proppant pack. Therefore, this unconfined proppant can be
re
easily washed out from the cell. After that, this situation repeats many times until all the
proppant is washed out completely or the pack stabilizes at height of 2.5 proppant grains. In this
lP

case, the proppant bridges the fracture walls and forms a very stable proppant pack even at
zero closure stress conditions.
na

The influence of closure stress on the proppant pack stability was clearly manifested in

showed that the bridging factor (h) is shifted to the range of greater // values, depending on
the experiments with the polymer adhesive tape. As it is shown later, experimental results
ur

stress. The pack stabilizes at h = 4.7 and 5.5 at 200 and 3000 psi of closure stress, respectively.
Jo

That value is close to 5-6 grains of the proppant arc reported in the literature (Asgian et al.,
1995; Milton-Tayler et al., 1992). The general view of the experimental points and fitting curves
for ISP 12/18 ceramic proppants are presented in the next section in Fig. 16.

After testing, the plate surfaces were visually inspected, and it was found out that after
each test, the proppant particles left deep imprints in the polymer tape (Fig. Error! Reference
source not found.) even at relatively low stress. That proves that embedment is a root-cause of
the pack stabilization even at relatively lowstresses. Hence, we may suggest the following
mechanism of the proppant pack stabilization. In addition to 2.5 layers of proppant that are
immobilized due to bridging, two more layers are immobilized due to embedment (one layer of
proppant grains that is adjacent to the bottom plate and one layer that is adjacent to the top
plate), and one to two more layers can be immobilized due to surface roughness.

18 | P a g e
of
Figure 10. Signs of proppant embedment in polymer adhesive tape after test with LWP 12/18 at 3000 psi.

ro
Proppant type. Milton-Tayler et al. (1992) observed that the coarser sintered proppant is
-p
more stable than fused proppant. The reason is the higher friction coefficient of rougher and
more angular material.
re
In Fig 15, we show the experimental data points obtained for fracturing sand (FS 20/40)
lP

and ceramic proppants (LWP 12/18 and 20/40) obtained in similar testing conditions. As it can
be seen from the graph, the fitting curve for FS 20/40 is slightly above the LWP 20/40 fitting
curve, which indicates that the sand pack is more stable than the pack of ceramics.
na

Fluid viscosity. In Milton-Tayler et al. (1992), viscosity was not specially studied as a
factor influencing the proppant flowback. Later, Parker et al. (1999) pointed out that fluid
ur

words, the viscosity decreases the fluid critical velocity. Indeed, turning from water (0.65 cP) to
viscosity increases drag force, which surely increases the risk of proppant flowback. In other

water-glycerol mixture (4.7 cP) in our experiments, the critical fluid velocity decreased by a
Jo

factor of three (from 1.5 to 0.5 cm/s) at // > 2.5 (Fig. 12). Note that such increase of fluid
viscosity by eight times did not compromise the pack stability at // < 2.5.

Unfortunately, due to equipment limitations, we were not able to do more extensive


research and test fluids with lower or higher viscosities, which would represent produced gas,
or heavy oil, as well as to do the tests with water-oil emulsions.

Geometry of proppant washing out. The theoretical considerations above make it clear
that proppant must start to flow out from the edge of a pack, which corresponds to the outlet of
a cell. Our experiments do not allow us to watch the process of proppant mobilization from the
beginning because of the nontransparent cell plates. However, after experiments, we observe
the through channels that formed in the proppant pack. For example, Fig. 7 shows a photo of
one particular geometry of a through-channel formed in the middle of a cell. In other tests, we
see that such channel is either attached to one of the edges of a proppant cell (Fig. 11, top left,
and bottom left) or not created at all (Fig. 11, top right). Observed channeling in proppant packs
is well-known from previous numerical studies of erosional instability of granular systems
(Cerasi and Mills, 1998; Parker et al., 1999). Aside from the instability, the position of a channel
in the tests, shown in Figs. 7 and 11, can be attributed to some perturbation of confining

19 | P a g e
stresses by slight angular deviation of confining plates. It is also interesting, that the proppant
can wash out from the edge of a pack and simultaneously redistribute the proppant mass, so
that after proppant flowback we still see homogeneous pack without any visible channels
(Fig. 11, top right).

of
ro
-p
re
lP
na
ur
Jo

Figure 11. Photos of proppant pack in a cell after the flowback (except bottom right, where the proppant has
kept stable) for LWP 20/40(top left), LWP 12/18 (top right) and ISP 12/18 (bottom left and right).

20 | P a g e
The bottom-right photo in Fig. 11 shows the distribution of proppant in a cell after the
test, where the proppant flowback has not been observed. The pack remained stable there even
at high flow rates, as we discussed in section 3.2.

4. THEORY VALIDATION BY EXPERIMENTS


4.1. Critical velocity for ceramic proppants
To validate the theory and criterion of proppant pack mobilization, we use the results of
laboratory experiments using the flowback apparatus described in the previous section. The
results are the dataset of measured critical filtration velocity that caused mobilization of
proppant of different types and mesh sizes: namely, the lightweight ceramic proppants (LWP) of
mesh sizes 12/18 and 20/40 and frac sand (FS) of mesh size 20/40. First, we compare the

of
model predictions for the experiments with steel plates only. Later, we proceed to the validation
of the theory for rough and soft fracture walls, where we use the series of experiments with

ro
polymer tape on the plates.

velocity ,D , fluid viscosity , flowback cell aperture and mean proppant diameter /. Other
-p
The following parameters were measured or known in the experiments: critical filtration

parameters, such as proppant permeability B, yield stress and bridging factor h were
re
unknown beforehand. However, we can evaluate some of them from the analysis of
lP

influence of proppant embedment and wall roughness on bridging , i.e., f ≪ /, g ≪ / in Eq. (31).
experimental results. For the experiments with steel confining plates, we can neglect the

Hence, according to the theory and previous observations, the bridging factor h = 3 is supposed
na

stress, B and , respectively, we introduce the so-called proppant stability parameter m and
to be between 2 and 3. To evaluate the unknown product of proppant permeability and yield
ur

rewrite the criterion (15) as follows:


m
,D = , m = 2B"
/ −h
/
Jo

(32)

where the parameter m has the physical dimension [Pa∙ m= ].


Every laboratory test has a measurement of the critical filtration velocity ,D given the
viscosity of fluid , proppant pack width , mean grain size /, and applied confining stress.
Criterion (32) allows us to evaluate the unknown proppant stability parameter m and bridging
factor h. Assuming that these parameters depend only on the type of proppant, we find them by
regression analysis of a large number of data points corresponding to the particular proppant.
The modified least squares method used for determination of the unknown parameters from
the experimental data points is described in Appendix A. Those data points that correspond to
the experiments, where the proppant was not mobilized, do not represent trusted values of
critical filtration velocity ,D , and so we exclude them from the regression analysis for the
parameter m. However, they were useful for evaluation of the bridging factor h, especially for
the experiments with the soft walls.
The first series of proppant flowback tests was done with LWP 12/18 proppant. In the

Such a wide range of // ratio allows us easily to observe the transition between the bridging
tests, we used different thicknesses of proppant pack, from 2 to 6 diameters of proppant grain.

21 | P a g e
( //~ 2 − 3) and mobilization at low filtration velocities ( //~5 − 6). To examine the
dependency of critical velocity on the confining stress, in different tests we applied as low stress

water ( = 0.65 cP) and a more viscous mixture of water and glycerol ( = 4.7 cP). Fig. 12
as 100 psi and as large value as 4000 psi. We also filtrated two types of fluids: less viscous tap

shows a summary of the data points for critical filtration velocity obtained in the laboratory
tests with LWP 12/18 and different proppant pack thickness, stresses and fluids. We compare
the laboratory data with the model predictions by criterion (32).

of
ro
-p
re
lP
na

Figure 12. Dependency of critical filtration velocity on proppant pack width in mean proppant diameters for

proppant with steel plates. Lines are the model (32) with parameters m = 2.2 ∙ 10>o p= ∙ q`,
different closure stresses (100, 1000, 4000 psi) and fluid viscosities (4.7, 0.65 cP) for LWP 12/18

h = 2.2. Points are the laboratory experiments.


ur

The model prediction shown in Fig. 12 is built after determination of the parameters m and h by
Jo

the regression analysis. They are found as m = 2.2 ∙ 10>o m= ∙ Pa and h = 2.2, respectively. The
obtained value of the stability parameter m helps us evaluate the possible value of solid pressure
at the edge region given the proppant pack permeability or vice versa.
In Fig. 12, we observe infinite growth of critical velocity at small values of // approaching the
bridging limit (~2.5) and their decay to small values far away from bridging, in accordance with
the theory predictions. The effect of filtrating fluid viscosity on proppant mobilization is also
well confirmed by the experimental results, which report lower value of the critical velocity at
larger viscosity of filtrating fluids. Finally, complete independence of the proppant mobilization
threshold on the applied confining stress in Fig. 12 supports our theoretical hypothesis about
primary role of the proppant pack edge in proppant mobilization, as far as edge stresses are
independent of the applied confining stress.
Next, a series of proppant flowback experiments was conducted for other types of
proppants (LWP 20/40 and FS 20/40). Similar to the previous tests, they were also fulfilled
with a wide range of pack thickness and confining stress and two types of filtrating fluids with
small and large viscosities. The corresponding results are shown in Figs. 13 and 14.

22 | P a g e
of
ro
-p
re
Figure 13. Dependency of critical filtration velocity on proppant pack width in mean proppant diameters for

LWP 20/40 proppant with steel plates. Lines are the model (32) with parameters , h = 2,
different closure stresses (100, 500, 1000, 2000, 3000 psi) and fluid viscosities (4.7, 0.65 cP) for
lP

m = 2 ∙ 10>o p= ∙ q`. Points are the laboratory experiments.


na
ur
Jo

Figure 14. Dependency of critical filtration velocity on proppant pack width in mean proppant diameters for

plates. Lines are the model (32) with parameters , h = 2, m = 1.95 ∙ 10>o p= ∙ q`. Points are the
different closure stresses (100, 2000, 3000 psi) and fluid viscosity 0.65 cP for FS 20/40 with steel

laboratory experiments.

these tests for different // ratio, fluid viscosities and stress conditions (Figs. 13, 14) leads to
Comparison of the theoretical criterion of mobilization and experimental data points in

23 | P a g e
the same conclusions that support the validity of the proppant pack mobilization theory that we
built in this work.
It is instructive to provide a comparative summary of the three test series for the three
types of proppant that we described above. For that purpose, we combine results of all
experiments with different proppants shown in Figs. 12-14 in one plot (Fig. 15). The data points
combine tests with different pack thickness and confining stress but only one selected filtrating

stability parameter m and bridging factor h that fit all three types proppants. Table 3
fluid. We also used the combined data set of measurements to calculate the values of proppant

summarizes these model parameters for three different types of proppants.

of
ro
-p
re
lP
na
ur

Figure 15. Dependency of critical filtration velocity on proppant pack width in mean proppant diameters for
Jo

different closure stresses (100-3000 psi) and fluid viscosity 0.65 cP for different types of ceramic

Lines are the model (32) with parameters , h = 2.2, m = 1.87 ∙ 10>o p= ∙ q`. Points are the
proppant: LWP with the mesh 12/18 and 20/40 and FS with the mesh 20/40, all with steel plates.

laboratory experiments.

LWP 12/18 LWP 20/40 FS 20/40 All Proppants

j, tt 1.408 0.74 0.567 1.03

u, vw ∙ tx 2.2 ∙ 10>o 2 ∙ 10>o 1.95 ∙ 10>o 1.87 ∙ 10>o

y 2.2 2 2 2.2
Table 3. Parameters of the model for critical filtration velocity obtained by the experimental data fitting
technique for three types of conventional ceramic proppants

In the last column of Table 3, we provide the common model parameters that fit all
experimental points for all used ceramic proppants (Fig. 15). These common model parameters

proppant and its mesh size. It is remarkable that the proppant stability parameter m, if
can be used to get an estimation of the critical filtration velocity for all types of ceramic

24 | P a g e
demonstrates, whilst their mean grain size / differs significantly. This interesting observation
measured precisely, varies insignificantly among different types of proppant, as Table 3

makes the criterion (32) almost universal for ceramic proppants of different mesh size.

4.2. Proppant pack cohesion assessment


The developed theory of proppant mobilization also lets us evaluate the cohesion of a
proppant pack sandwiched between two walls. Cohesion is the shear strength of the material
that needs to be overcome to initiate shear motion (Howard et al., 1999; Richefeu et al., 2006;

is rarely measured in laboratory. Referring to the definition of proppant stability parameter m in


Than et al., 2017). This is the intrinsic property of any granular suspension in dense state, which

permeability B:
(32), we obtain the following dependence of the yield stress on the proppant pack

of
=
2B"
(33)

ro
This expression gives us a clue to estimate yield stress that holds the pack at the edge stable at
-p
critical filtration velocity and estimated the value of stability parameter m and proppant pack
low filtration velocities. It enables the yield stress estimation provided we have measured the

permeability B at low confining stress from the experiments.


re
lP

The cohesion of proppant can be evaluated from the yield stress using the following
formula for the uniaxial compressive strength (Jaeger et al., 2007):

= 2 z 1 + "= {/=
+" |
na

(34)

where " is the coefficient of interparticle friction. Taking the value of friction coefficient
ur

" = 0.32 from the experiments with PMMA (poly(methyl methacrylate)) spheres (Boyer et al.,
Jo

2011), from (34) we have

= 0.365 (35)

In Table 4, for the proppants LWP 12/18, 20/40 and FS 20/40, we calculate the values of

stability parameter m and permeability B experimentally measured for those proppants at low
proppant yield stress and cohesion using (33)-(34). For that, we use the obtained values of

confining stress (100 psi).

LWP 12/18 LWP 20/40 FS 20/40

j, tt 1.408 0.74 0.567

}, ~w•€• 2700 650 220

) , vw 14 52 150

‚) , vw 5.1 18.9 54.8

25 | P a g e
Table 4. Comparison study of different parameters at the proppant pack edge: proppant pack permeability B,
yield stress , and cohesion of the proppant pack .

We do not measure the value of proppant pack cohesion directly in this work. However, it
is of interest to compare the calculated values with existing measurements of granular medium
cohesion. For example, Richefeu et al. (2006) experimentally investigated shear strength
properties of wet granular materials for natural sand and glass beads. The main difference with
our experiments is that authors in this work consider unsaturated granulates for which
cohesion forces are determined by capillary forces between the contiguous solid grains.
However, the cohesion force is also present owing to a granular structure of solid particles in a
pack confined by rigid walls. One more mechanism is adhesion, which may exist between the
particles as, for example, in resin-coated proppants. These are the different mechanisms that
can hold a pack of proppant particles stable in the edge region when the filtration velocity does
not exceed its critical value.

of
We presume that in our experiments capillary forces may also be present due to the
uneven structure of the proppant pack, i.e., there may appear “empty” regions with a gas phase,

ro
forming liquid bridges at the contact points and giving rise to finite cohesion. The values of

monodisperse glass beads (/ = 1 mm) to 600 Pa for sands (/ = 0.16 mm). As expected, the
cohesion obtained in the experiments after Richefeu et al. (2006) are in the range of 150 Pa for
-p
cohesion force in these experiments is higher than that estimated in the Table 4 from our
re
experiments. This can be explained by the presence of capillary forces and less regular shapes of
particles in the experiments by Richefeu et al. (2006). Nevertheless, the cohesion magnitudes in
lP

both types of experiments are much closer to each other than if they are compared to the typical
magnitude of confining stresses applied to proppant pack, which are several orders larger. That
proves our initial hypothesis that the resistance to proppant pack mobilization is created by the
na

cohesion of weakly stressed proppant at the edge of a pack.


ur

4.3. Stress-dependent bridging


Jo

the walls covered with polymer tape. We observed that the bridging factor, i.e., the critical //
Another series of proppant flowback experiments was conducted in the flowback cell with

ratio, at which the proppant pack becomes stable even at very high flow rates, is increased when
the softer walls are used. For each proppant type, we ran a series of experiments at different
confining stress. We noted that the bridging factor is increasing with the increase of confining
stress . This dependence is a result of proppant embedment into cell walls, that we describe
in the theory (31) and illustrate in the experimental section (Fig. 16).
In Fig. 16 we show how the critical velocity depends on the ratio of proppant pack width
and mean proppant diameter taking the example of ISP proppant with the mesh 12/18. The
open circles in Fig. 16 should not be regarded as a critical flowback velocity because they
correspond to the experiments in which the proppant pack was not mobilized. We have put

bridging factor h is defined as the midpoint between the minimum // ratio with unstable
them here only to show when bridging already took place. For each particular stress, the

proppant pack and the maximum // ratio, where the proppant is still stable. One can see that
the model curves reproduce the results of experiments with sufficient accuracy.

26 | P a g e
of
ro
different closure stresses ( = 200, 500, 1000, 3000 ƒ„) and fluid viscosity ( = 4.67 ∙
Figure 16. Dependency of critical flowback velocity on fracture width in mean proppant diameters for

10>N q` ∙ ƒ) for ISP proppant with the mesh 12/18 in experiments with rough cell walls. Lines
-p
correspond to the model (32) with parameters, m = 4.4 ∙ 10>o p= ∙ q`; bridging parameters for
re
different stresses were set manually. Points correspond to laboratory experiments. Open circles
correspond to the situations in which there was no flowback (FB) observed and proppant pack
lP

remained stable.

Fig. 17. illustrates how bridging factor h depends on the confining stress according to
na

the experiments with the ISP proppant shown in Fig. 16. We suppose that the actual value of
bridging factor for each stress is within the error bars indicated in Fig. 17.
ur
Jo

12/18. Error bars correspond to the limiting points, where the upper boundary is a minimum //
Figure 17. Dependency of bridging factor on confining stress, observed in experiments with ISP with the mesh

27 | P a g e
ratio with unstable proppant pack and the lower boundary is a maximum // ratio where the
proppant is still stable.

Here we do not provide an analytical model for the bridging factor h as a function of
confining stress because this dependency can be different for other materials and rocks.

bridging factor h can be determined.


However, the same measurement procedure can be fulfilled for any given rock type, and the

CONCLUSIONS
Despite many studies on proppant flowback in the literature over the last 30 to 40 years,
which have improved our understanding of the phenomenon, there is still a need for a
mathematical quantitative criterion of proppant pack mobilization that can be readily
integrated into existing simulators. In this work, we derived and validated such a quantitative

of
criterion. It is written in terms of the minimum filtration velocity that triggers proppant flow in

ro
a fracture and is sensitive to several important factors that are known to affect proppant pack
mobilization. They are fracture width (w), ratio of fracture width to the mean proppant
-p
diameter (d), proppant pack cohesion ( ), and proppant pack permeability (B) and filtrating
fluid viscosity ( ).
re
The offered criterion describes the well-known effect of bridging, when the proppant pack

First is reaching the minimum number of particles per width of a pack, i.e. , //, which hinders
stays immobile even at high filtration velocities. There are three factors that cause bridging.
lP

shear flow of particles. We evaluated that minimum to be 2.5. Second is the roughness of
na

fracture walls, which makes part of proppant pack adjacent to walls immobile due to their
nonflatness. The third factor is the softness of fracture walls, which facilitates embedment of
particles into wall, and also creates effective reduction of proppant pack width.
ur

We also recognized that leading role of a loosely stressed edge of a proppant pack in
Jo

mobilizing the pack. The stress at the edge is defined by the cohesive properties of a granular
pack and is completely independent of the applied confining stress. It explains our experimental

laboratory experiments showed, though, that the applied stress influences the bridging factor h,
evidence that the critical filtration velocity is independent of the confining stress. Our

if the walls are soft, so that the larger the stress, the higher the bridging factor is.
As mentioned, the edge stresses are proportional to the cohesion of a proppant pack. The
cohesion determines critical filtration velocity for proppant pack mobilization. It particularly
explains why the proppants with larger cohesion, such as the resin-coated ones, are more
stable. It is also remarkable to note that even bare ceramic proppants, which we experimentally
examined in this work, appear to have a final cohesion. Our estimations for the magnitude of
cohesion roughly agrees with the direct measurements of cohesion of wet sands, which are all
estimated by order of hundreds of pascals, and depend on grain sphericity, surface, and
capillary forces. The cohesion and critical velocity of proppant flowback are thus coupled. One
can either define the cohesion of the given granular material and calculate the critical velocity or
measure the critical velocity and determine the cohesion of a pack.
In this work, we also conducted a large series of laboratory tests that mimic fluid filtration
through a real proppant packs within a fracture under subterranean stress conditions. In the

28 | P a g e
tests, we used different types of commercial proppants and typical fluid viscosities and filtration

several centimeters per second, unless // ratio is close to the bridging factor. Experiments
rates and observed that the filtration velocity to cause proppant mobilization is of order of

indicated that onset of proppant flowback is always associated with a step-wise drop of the
proppant pack height in a cell.
The experimental measurements of critical filtration velocity allowed us to validate the
theoretical criterion at different proppant pack widths, mesh sizes and fluid viscosities and with
stiff and soft walls. Using a fitting procedure, we found the unknown parameters of proppant
stability for different types of proppants. We managed to show that for a family of ceramic
proppants, there exists a similarity of this stability parameter for different mesh sizes of
proppants. That gave us a clue to obtain a unique value of the proppant stability parameter
applicable to a whole family of ceramic proppants of different mesh size. We suggest using this
unique parameter in the criterion of proppant mobilization.
Experiments allowed us to detect the dependency of bridging factor h on the confining

of
stress as a result of proppant particle embedment into soft walls. We also experimentally

ro
demonstrated a variety of possible scenarios of proppant washing out in a laboratory cell during
fluid filtration. The latter may be used by future researchers to explain those behaviors
theoretically and numerically. -p
Results of this work can be directly used in the proppant flowback simulators and applied
re
to various proppant flowback mitigation workflows (Chuprakov et al., 2020). For dealing with
other types of proppants that we did not cover in this work, we leave readers a possibility to
lP

determine their cohesion properties using a similar technique.


na

ACKNOWLEDGEMENTS
ur

The authors are thankful to Denis Bannikov, Vadim Isaev, Maxim Ivanov, Dean Willberg,
Pavel Spesivtsev, Maxim Chertov, Stepan Yukhtarov, Vladimir Radaev, Ludmila Belyakova and
Jo

Vladimir Abashkin from Schlumberger for technical review of this work, as well as
Schlumberger for permission to publish it.

REFERENCES
Asgian, M.I., Cundall, P.A., Brady, B.H., 1995. The Mechanical Stability of Propped Hydraulic
Fractures: A Numerical Study. J. Pet. Technol. 47, 203–208.
https://doi.org/10.2118/28510-PA
Boyer, F., Guazzelli, E., Pouliquen, O., 2011. Unifying suspension and granular rheology. Phys Rev
Lett 107, 188301. https://doi.org/10.1103/PhysRevLett.107.188301
Canon, J.M., Romero, D.J., Pham, T.T., Valko, P.P., 2003. Avoiding Proppant Flowback in Tight-Gas
Completions with Improved Fracture Design. https://doi.org/10.2118/84310-MS
Cerasi, P., Mills, P., 1998. Insights in erosion instabilities in nonconsolidated porous media. Phys.
Rev. E - Stat. Physics, Plasmas, Fluids, Relat. Interdiscip. Top. 58, 6051–6060.
https://doi.org/10.1103/PhysRevE.58.6051

29 | P a g e
Chuprakov, D., Belyakova, L., Iuldasheva, A., Alekseev, A., Syresin, D., Chertov, M., Spesivtsev, P.,
Salazar Suarez, F.I., Velikanov, I., Semin, L., Bannikov, D., 2020. Proppant Flowback: Can We
Mitigate the Risk? SPE Hydraul. Fract. Technol. Conf. Exhib.
https://doi.org/10.2118/199748-MS
Dontsov, E. V., Peirce, A.P., 2014. Slurry flow, gravitational settling and a proppant transport
model for hydraulic fractures. J. Fluid Mech. 760, 567–590.
https://doi.org/10.1017/jfm.2014.606
Elliott, S.J., Gale, J.F.W., 2018. Analysis and distribution of proppant recovered from fracture
faces in the HFTS slant core drilled through a stimulated reservoir. SPE/AAPG/SEG
Unconv. Resour. Technol. Conf. 2018, URTC 2018. https://doi.org/10.15530/urtec-2018-
2902629
Fjaer, E., Holt, R.M., Horsrud, P., Raaen, A.M., Risnes, R., 2008. Petroleum related rock mechanics,
in: Petroleum Related Rock Mechanics. Elsevier.
Garagash, I.A., Osiptsov, A.A., Boronin, S.A., 2019. Dynamic bridging of proppant particles in a

of
hydraulic fracture. Int. J. Eng. Sci. 135, 86–101.
https://doi.org/10.1016/j.ijengsci.2018.11.004

ro
Ghanizadeh, A., Clarkson, C.R., Deglint, H., Vahedian, A., Aquino, S., Wood, J.M., 2016.
-p
Unpropped/Propped Fracture Permeability and Proppant Embedment Evaluation: A
Rigorous Core-Analysis/Imaging Methodology. https://doi.org/10.15530/URTEC-2016-
2459818
re
Guazzelli, É., Pouliquen, O., 2018. Rheology of dense granular suspensions. J. Fluid Mech. 852,
lP

P11–P173. https://doi.org/10.1017/jfm.2018.548
Guo, B., Ugwu, D.O., 2015. An Experimental Investigation of the Critical Flowback Velocity in
Hydraulic-Fracturing Shale Gas Wells. Hydraul. Fract. J. 2, 19–25.
na

Howard, P.R., James, S.G., Milton-Tayler, D., Willcms-Braun, B., 1999. High-Permeability
Channels in Proppant Packs Containing Random Fibers. SPE Prod. Facil. 14, 197–202.
ur

https://doi.org/10.2118/57392-PA
Jaeger, J.C., Cook, N.G.W., Zimmermann, R.W., 2007. Fundamentals of Rock Mechanics, 4th ed,
Jo

Rock Mechanics. Blackwell Publishing. https://doi.org/10.1016/0040-1951(77)90223-2


Lecampion, B., Garagash, D.I., 2014. Confined flow of suspensions modelled by a frictional
rheology. J. Fluid Mech. 759, 197–235. https://doi.org/10.1017/jfm.2014.557
Li, K., Gao, Y., Lyu, Y., Wang, M., 2015. New Mathematical Models for Calculating Proppant
Embedment and Fracture Conductivity. SPE J. 20, 496–507.
https://doi.org/10.2118/155954-PA
McLennan, J., Walton, I., Moore, J., Brinton, D., Lund, J., 2015. Proppant backflow: Mechanical and
flow considerations. Geothermics 57, 224–237.
https://doi.org/10.1016/j.geothermics.2015.06.006
Milton-Tayler, D., Stephenson, C., Asgian, M.I., 1992. Factors Affecting the Stability of Proppant
in Propped Fractures: Results of a Laboratory Study. SPE Annu. Techncial Conf. Exhib.
https://doi.org/10.2118/24821-MS
Morris, J.P., Chugunov, N., Meouchy, G., 2014. Understanding Heterogeneously Propped
Hydraulic Fractures Through Combined Fluid Mechanics, Geomechanics, and Statistical
Analysis. 48th US Rock Mech. / Geomech. Symp. 2014.
Parker, M., Weaver, J., Van Batenburg, D., 1999. Understanding Proppant Flowback. SPE Annu.

30 | P a g e
Tech. Conf. Exhib. https://doi.org/10.2118/56726-MS
Pearson, J.R.A., Zazovsky, A.F., 1997. A Model for the Transport of Sand Grains From a
Perforation During Underbalance Surge. SPE Annu. Techncial Conf. Exhib.
https://doi.org/10.2118/38634-MS
Potapenko, D., Theuveny, B., Williams, R., Moncada, K., Campos, M., Spesivtsev, P., 2019. State of
the Art of Flow Management for Frac Plug Drillout and Flowback Proppant Flowback
Impact. SPE Annu. Tech. Conf. Exhib. https://doi.org/10.2118/196084-MS
Potapenko, D.I., Williams, R.D., Desroches, J., Enkababian, P., Theuveny, B., Willberg, D.M.,
Moncada, K., Deslandes, P., Wilson, N., Neaton, R., Mikovich, M., Han, Y., Conort, G., 2017.
Securing Long-Term Well Productivity of Horizontal Wells Through Optimization of
Postfracturing Operations. SPE Annu. Tech. Conf. Exhib. https://doi.org/10.2118/187104-
MS
Pyrak-Nolte, L.J., Morris, J.P., 2000. Single fractures under normal stress: The relation between
fracture specific stiffness and fluid flow. Int. J. Rock Mech. Min. Sci. 37, 245–262.

of
https://doi.org/10.1016/S1365-1609(99)00104-5

ro
Richefeu, V., El Youssoufi, M., Radjaï, F., 2006. Shear strength properties of wet granular
materials. Phys. Rev. E 73, 51304. https://doi.org/10.1103/PhysRevE.73.051304
-p
Shor, R.J., Sharma, M.M., 2014. Reducing Proppant Flowback From Fractures: Factors Affecting
the Maximum Flowback Rate. SPE Hydraul. Fract. Conf. https://doi.org/10.2118/168649-
re
MS
Than, V. Du, Khamseh, S., Tang, A.M., Pereira, J.-M., Chevoir, F., Roux, J.-N., 2017. Basic
lP

Mechanical Properties of Wet Granular Materials: A DEM Study. J. Eng. Mech.


https://doi.org/10.1061/(ASCE)EM.1943-7889.0001043
na

Velikanov, I., Isaev, V., Bannikov, D., Tikhonov, A., Semin, L., Belyakova, L., Kuznetsov, D., 2018.
New Fracture Hydrodynamics and In-Situ Kinetics Model Supports Comprehensive
Hydraulic Fracture Simulation. SPE Eur. Featur. 80th EAGE Conf. Exhib.
ur

https://doi.org/10.2118/190760-MS
Zhang, Z., Guo, B., 2016. The Critical Flow back Velocity in Hydraulic-Fracturing Shale Gas Wells.
Jo

Int. J. Eng. Res. Appl. 6, 7–11.

31 | P a g e
METRIC

psi × 6.894 E+03 = Pa


cP E-03 = Pa ∙ s
Darcy E-12 = m=

APPENDIX A. METHOD OF MODEL PARAMETER FITTING FROM


THE EXPERIMENTAL DATA

of
Consider the following functional † that depends on two parameters m and h:

m >{ =

ro
† m, h = ‡ ˆ − −h Œ ,
WŠ‹ ‰
(A-1)

,‰ / /
,D,‰
‰Ž{
-p
where • is the number of performed experiments for the particular proppant with the mean
diameter of grains /, ,D,‰ is the measured value of critical velocity in the „ •‘ experiment and
WŠ‹
re
,‰ ‰
lP

and are the corresponding values of the fluid viscosity and cell width.

So, we want to find the parameters m and h, which minimize the functional (A-1). For that,
we create a mesh for the parameter h in the range ’2,3“ (these values were observed in the
experiments with smooth cell walls) with the step Δh, so that
na

2 3
h• = B ∙ Δh, B
Δh Δh
(A-2)
ur

In our calculations, we used Δh = 10>{ . For each h• , we find the value of parameter m, which
Jo

minimizes the functional †• m = † m, h• :



m >{ =
†• m = ‡ ˆ − − h• Œ → min
WŠ‹ ‰
(A-3)

,‰ / /
,D,‰
‰Ž{

The corresponding value of m, minimizing (A-3), we denote as m• , which can be found by solving
the following equation:
/†• m
=0
/m
(A-4)

Substituting (A-3) into (A-4), we obtain



/†• m• m• >{ 1 >{
= −2 ‡ ˆ − − h• Œ ∙ − h• =0
WŠ‹ ‰ ‰
(A-5)
/m ,D,‰
,‰ / / ,‰ / /
‰Ž{

Unifying the terms, containing m• , we obtain the following expression:

32 | P a g e
• • WŠ‹
1 >{ = >{
m• ‡ ˆ − h• Œ =‡ − h•
‰ ,D,‰ ‰
(A-6)

,‰ / / ,‰ / /
‰Ž{ ‰Ž{

So, the value of m• is from (A-6) as


WŠ‹
>{
∑‰Ž{ ,D,‰ ‰
− h•
,‰ / /
(A-7)

m• = >= ,
1
∑‰Ž{ ‰
− h•
,‰ /
= /

So, at each k-th step along the mesh (2⁄Δh B 3⁄Δh), we get the arguments (m• , h• )
and value of functional † m• , h• = †• . The value of k corresponding to the minimum of †• gives
us the sought values of model parameters m• , h• .

the additional mesh must be introduced for other non-linear parameters like h.
This method can be easily generalized for the models with multiple parameters, where

of
ro
-p
re
lP
na
ur
Jo

33 | P a g e
Criterion of proppant pack mobilization by
filtrating fluids: theory and experiments
Dimitry Chuprakov1, Aliia Iuldasheva2
Schlumberger Moscow Research, 13 Pudovkina St., Moscow, Russia, 119285
Alexey Alekseev3
Schlumberger Novosibirsk Technology Center, 1/10 Zelyonaya Gorka,
Novosibirsk, Russia, 630060

of
HIGHLIGHTS

ro
• Critical filtration velocity for proppant pack mobilization is derived and validated.

-p
Mobilization of a compressed pack occurs at the edge and independent of the stress.

re
Pack stability depends on particle size and cohesion, wall roughness and softness.
• Criterion is unique for a whole family of ceramic proppants of different mesh size.
lP

• Possible geometries and scenarios of proppant washing out revealed experimentally.


na
ur
Jo

1
Corresponding author, email: dchuprakov@slb.com
2
Email: agazizova@slb.com
3 Email: AAlekseev4@slb.com

1|Page
Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

of
ro
-p
re
lP
na
ur
Jo

Schlumberger-Private

You might also like