You are on page 1of 18

FEATURE ARTICLE

Multifunctional Flocculants www.mrc-journal.de

Advanced Polymer Flocculants for Solid–Liquid Separation


in Oil Sands Tailings
Sarang P. Gumfekar, Vahid Vajihinejad, and João B. P. Soares*

volume of tailings is projected to reach


The generation of tailings as a by product of the bitumen extraction process two billion cubic meters in 2034 if tail-
is one of the largest environmental footprints of oil sands operations. Most ings management procedures remain the
same.[3]
of the tailings treatment technologies use polymer flocculants to induce
When fresh tailings consolidate for a
solid–liquid separation. However, due to the complex composition of tailings, few years after their transfer to tailings
conventional flocculants cannot reach the same performance achieved in ponds, gravitational forces cause the coarse
other wastewater treatments. Over the last couple of decades, the oil sands components (sands) to settle, releasing a
industry has used acrylamide-based flocculants to treat tailings, achieving limited amount of the water to the surface.
major progress in process optimization and integration with mechanical The bottom part of the remaining suspen-
sion is called mature fine tailings (MFT).
operations, but they still could not reach the required land reclamation Typically, MFT contains negatively charged
targets. Over the last 5 years, the group designed, synthesized, and tested clays (≈30–35 wt%), water (≈65 wt%), and
several novel polymer flocculants tailored for oil sands tailings treatment. residual bitumen (3–5 wt%).[4] Compared
This feature article communicates recent developments in these innova- to fresh tailings, which still contain coarse
tive polymers. The article first provides a background on tailings generation particles that consolidate over time, MFT
is a gel-like suspension that consolidates
and treatment, followed by the description of advanced polymer flocculants
extremely slowly. Clays in MFT con-
categorized according to their microstructures such as linear, branched, and tain mainly kaolinite, illite, and mixed-
graft. The other tailings remediation technologies and one of the initial works layer clays such as interstratified illite–
on modeling of tailings flocculation is discussed. smectite and kaolinite–smectite.[5] Fine
clays are mostly responsible for the water
retained in MFT due to their high sur-
face area. Additionally, fugitive bitumen
1. Introduction binds clay particles, constricting interparticle pathways and
trapping water.
1.1. Generation of Oil Sands Tailings

Oil sands deposits in Canada are the third largest oil reserve 1.2. Colloidal Stability of Tailings and Challenges
in the world, with deposits estimated at 28.3 billion cubic in Dewatering
meters.[1] In the Clark hot water bitumen extraction process, oil
sand ores are mined, crushed, and mixed with warm alkaline Since MFT contains a mixture of negatively charged clays, it
water, followed by the separation of bitumen using a flotation shows strong negative zeta potential that causes colloidal sta-
process.[2] Figure 1 shows the schematic of oil sands extraction bilization. Generally, clays contain hydroxyl and oxide func-
emphasizing the main elements of the process. The extraction tional groups that make their particles negatively charged at
of bitumen from oil sands using the process shown in Figure 1 neutral pH (pH = 7) and even more negative as pH increases.
forms waste tailings that cause technical and environmental Thus, the stability of clays in MFT increases with increase in
problems when discharged into tailing ponds. Currently, the pH. Current bitumen extraction processes use caustic addi-
tailing ponds in Canada have polluted more than 200 km2, tives to liberate bitumen from oil sands, generating alkaline
which is approximately equivalent to 111 000 hockey rinks. The tailings with pH of approximately 8.5. This alkaline medium
increases the electrostatic repulsive forces between clays
and significantly increases MFT stability. Surface charges
Dr. S. P. Gumfekar, V. Vajihinejad, Prof. J. B. P. Soares also arise when ions from the neighboring solution adsorb
Department of Chemical and Materials Engineering on an initially neutral solid surface. Another source of sur-
University of Alberta face charges is the dissociation of surface groups. The silica
Edmonton, AB T6G 2V4, Canada
E-mail: jsoares@ualberta.ca
surface of clays can react with water to form silicic acid,
which can then dissociate to form a silicic anion, as shown
The ORCID identification number(s) for the author(s) of this article
can be found under https://doi.org/10.1002/marc.201800644. below.

DOI: 10.1002/marc.201800644 (SiO2 )S + H2O ↔ (H2SiO3 )S ↔ (HSiO3− )S + H+ (1)

Macromol. Rapid Commun. 2019, 40, 1800644 1800644  (1 of 18) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mrc-journal.de

As the reaction shown in Equation (1) proceeds in alkaline pH,


negative charges build up on the solid surface and H+ ions dif- Sarang P. Gumfekar is a
fuse into the bulk solution. The negative charges generated on postdoctoral fellow at the
clay surfaces via this mechanism induce significant repulsive University of Alberta, Canada.
forces to make them colloidally stable. His work focuses on the
Polymer flocculants and coagulants reduce the repulsive design of new polymers and
forces via charge neutralization and bridging. The current control of their microstruc-
polymer-based technologies, however, struggle to meet the fol- tures. He obtained his Ph.D.
lowing multiple requirements for effective MFT treatment: in chemical engineering from
the University of Alberta,
i. Settling: The extremely slow settling rate of MFT is attributed
under the supervision of Prof.
to the use of caustic additives during the bitumen extraction
João Soares. He also received
process that gels ultrafine particles (50–400 nm) in MFT, in-
an M.A.Sc. degree from the
creasing its viscosity and making it difficult to consolidate.
University of Waterloo, Canada, and a B.Eng. degree from
The surface charge of the fine particles is also responsible for
the University of Pune, India, both in chemical engineering.
stabilizing the MFT suspension.
Alongside polymer synthesis, he is interested in physico-
ii. Dewaterability: MFT holds a significant amount of water at
chemical processes such as flocculation, nanofiltration,
alkaline pH. Commonly employed polymer flocculants are
and advanced oxidation processes for water treatment.
usually water soluble and aggravate this problem, trapping
water even after floc formation. The presence of residual bi- Vahid Vajihinejad is a
tumen on the surface of clays further limits the adsorption Ph.D. candidate at the
of conventional water-soluble flocculants, making porous and University of Alberta, Canada,
fragile aggregates holding a large amount of water. under the supervision of
iii. Quality of recovered water: The water recovered from tailings Prof. João Soares. His
should be recycled back to the bitumen extraction process, research focuses on studying
but its quality affects the efficiency of bitumen extraction polymer microstructure–
from oil sands. Although charge neutralization of clays using flocculation performance rela-
inorganic cations such as Ca2+ or Al3+ may seem to be an easy tionships, flocculation mode-
solution for tailings consolidation by coagulation of particles, ling, and reaction engineering
these cations are detrimental to the bitumen extraction pro- of water-soluble polymers.
cess as it reduces the recovery and quality of the froth by a He obtained his M.E.Sc. in
mechanism known as slime coating. Presence of even small chemical engineering from Western University, Canada, and
amount of fines in the supernatant (turbidity) causes damage B.Sc. from the Sharif University of Technology, Iran.
to the process equipment.
iv. Solids content of the sediments: Current polymer flocculants João B. P. Soares is a
can treat MFT to a certain extent, but MFT needs to be di- professor in the Department
luted to promote effective interactions between polymer and of Chemical and Materials
clay particles. Thus, it is imperative to remove both internal Engineering at the University
and externally-added water from MFT to produce trafficable of Alberta, Canada. He is a
sediments. Additionally, the water entrapped in the flocs re- Campus Alberta Innovates
duces the solids content of the sediments, requiring further Program Chair in Interfacial
expensive mechanical treatments such as centrifugation and Polymer Engineering for
filtration. Oil Sands Processing, and
a Tier I Canada Research
1.3. Conventional Technologies Chair in Advanced Polymer
Reaction Engineering. His
The major methods used to treat MFT include the use of lime, main research interests are polymerization reaction
inorganic coagulants, polymer flocculants, pH control, and engineering, polyolefin synthesis and characterization, and
freeze–thaw cycles. The following section outlines two of the polymer flocculation engineering.
industrial methods that are currently being used. Consolidated
tailings and paste technology are the two technologies adopted
by the industry.
The consolidate tailings method involves mixing densified
extraction tailings (cyclone underflow) and MFT with a coagu- divalent ions that negatively affect the bitumen extraction pro-
lant (typically gypsum) to produce non-segregating tailings, cess.[6] Moreover, H2S may be released after treatment due to
which when discharged into a tailings pond form a rapidly the anaerobic reduction of SO42− with the residual bitumen in
consolidating deposit. This approach, however, suffers from a the tailings.[7]
few drawbacks. Gypsum significantly changes the chemistry Paste technology involves the settling of suspended fines
of the recycled water by imparting a high concentration of within a process vessel (thickener) promoted by polymer

Macromol. Rapid Commun. 2019, 40, 1800644 1800644  (2 of 18) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mrc-journal.de

Figure 1.  Schematic of the oil sands extraction process.

flocculants. Depending on the type of polymer, the flocculated also been used to quantify floc formation in tailings. Since the
fines settle quickly and form a stable solids bed that can be subsequent sections in this review describe the performance of
deposited in tailings ponds for land reclamation, allowing the recently synthesized polymer flocculants, it is important to first
supernatant water to be recycled to the bitumen extraction pro- describe how their efficiency is assessed.
cess. The choice of the flocculant and the associated parameters Settling rate measures how quickly a flocculant can settle
such as dosage and mixing conditions determine the effec- solids in suspension by tracking the solid–liquid interface
tiveness of paste technology. Although promising, paste tech- (mudline) as a function of time, forming a settling profile,
nology does not provide yet an effective solution for the existing such as the one shown in Figure 2A. Often, the initial settling
MFT inventory (legacy tailings) because they are convention- rate (ISR), the slope of the initial linear section of the settling
ally designed to handle thickening of low solids effluent. It is profile, is reported instead of the average settling rate. ISR is a
also capital-intensive since it requires design, construction, simple measurement of the initial flocculation kinetics.
and maintenance of thickener, and purchase of state-of-the-art CST quantifies how easily suspensions or sediments can
polymer flocculants. be dewatered. This technique can be used immediately after
Even though current technologies for sewage water treat- the polymer is mixed with tailings or after the sediments are
ment, dewatering of clay suspensions, removal of the organic recovered by removing the supernatant. Figure 2B shows an
content in water, and treatment of ceramic wastewater are image of a CST instrument, which measures the time required
mature, this is far from the truth for oil sands tailings reme- for water to move through a sheet of filter paper positioned
diation: no technology can effectively treat oil sands tailings between the two electrodes. A higher CST value indicates lower
and eliminate tailings ponds. Although acrylamide-based floc- dewaterability (high water retention) of the suspension or sedi-
culants are widely used in the industry, the sediments formed ments. Indirectly, it reflects the aggregates network formed
after flocculation still hold a significant amount of water. This by the polymer flocculant, an indication of how fast the inter-
knowledge gap calls for innovative multifunctional flocculants aggregate water can drain.
that can effectively flocculate MFT clays, form
sediments with high solids contents, and
release recycle water with adequate chem-
istry. This article describes some of the new
polymer flocculants developed in our group
to treat MFT. We also discuss remediation
strategies that are not entirely based on floc-
culation. Recent results on the modeling of
floc formation in oil sands tailings using in-
house flocculants are also discussed.

2. Performance Matrix for


Flocculant Evaluation
Flocculant performance is evaluated with
a few measurements such as settling rate,
supernatant turbidity, capillary suction time
Figure 2.  A) Schematic of a settling rate experiment and calculation of the initial settling rate
(CST), specific resistance to filtration (SRF), (ISR). Reproduced with permission.[8] Copyright 2015, American Chemical Society. B) Capillary
and solids content. More recently, focused suction time unit (CST, model 319 Triton Electronics). Reproduced with permission.[9] Copy-
beam reflectance measurement (FBRM) has right 2005, American Chemical Society.

Macromol. Rapid Commun. 2019, 40, 1800644 1800644  (3 of 18) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mrc-journal.de

Supernatant turbidity measures the fraction of fine particles FBRM is used to measure aggregate size without the need
that remained suspended in the supernatant after flocculation. for sampling and dilution. Figure  3 shows the working prin-
Often turbidity is measured using turbidimeter and expressed ciples of FBRM. The instrument measures the chord length
in Nephelometric Turbidity Unit (NTU). Some researchers have and also provides the real-time distribution of aggregates. It
used UV spectrophotometers to monitor the transmittance of is important to note that FBRM measures chord length and
the solution, but this technique is limited by a lower upper limit counts, but not number and size of aggregates. One can, how-
for the concentration of suspended particles.[10] Although sev- ever, correlate chord length with actual particle size measured
eral researchers have developed empirical correlations between by other techniques such as laser diffraction spectroscopy.[12]
turbidity and volume fraction of particles,[11] turbidity alone Recently, some researchers have used surface force measure-
provides a quick and reliable way to evaluate how efficiently the ments to indirectly assess the effectiveness of a flocculant in
flocculant can capture fine particles. terms of its adsorption on clay surfaces. Typically, surface force
The term specific resistance to filtration or filtration resist- apparatus (SFA) or atomic force microscopy (AFM) are used
ance is derived from Darcy’s law. It is a purely hydrodynamic to obtain force–distance curves. Lu et al. synthesized poly(2-
concept. The instrument only measures the volume or mass lactobionamidoethyl methacrylamide) of various molecular
of water released from polymer-treated tailings as a function of weights to dewater kaolin suspensions, a major component of
time, t. Then, this data is used in Equation (2) to calculate the tailings.[14] They studied the effect of molecular weight on the
SRF. interaction of polymers and kaolin surfaces using SFA. Poly-
mers with molecular weights of 83, 186, and 456 kDa had adhe-
2 b P A2 sion forces of 3, 8, and 11 mN m−1, respectively. These results
SRF = (2) indicated that higher molecular weight polymers, having more
µω
hydroxyl groups per chain, adhered more strongly to the sub-
where A is the filter area (m2), P is the applied pressure (Pa), strate via hydrogen bonding. The authors also verified these
µ is the viscosity of the filtrate (Pa·s), ω is the mass of solid results using conventional settling tests.
cake formed per unit filtrate volume (kg m−3), and b is the slope
determined from t/V against V curve. SRF quantifies the filter-
ability of a suspension. A small SRF value indicates that the 3. Multifunctional Linear Polymer Flocculants
filter cake is highly permeable. SRF and the mass of the solid
cake (ω) are inversely proportional: the higher the mass of the The most common flocculants used in the oil sands industry
solid cake, the lower the SRF. are copolymers of acrylamide and acrylic acid known as anionic
One of the objectives of polymer-assisted flocculation of tail- polyacrylamides (A-PAM), (Table  1) typically having molecular
ings is to remove as much water as possible, thus increasing weight averages of several million.[15] During flocculation, the
the solid contents of the sediments. Solid contents are usually polymer may be applied directly to the MFT or after destabi-
measured by a simple gravimetric method, wherein a sample lizing it via coagulation.[16] Prior coagulation with divalent
of known mass is heated until all liquid is evaporated. The dif- cations helps overcome repulsive forces between negatively
ference between initial and final masses is used to calculate charged clay particles, giving access to bridging flocculation
the solids content. Since often tailings are diluted to allow for and reducing the required dosage of the polymer.[17] A system-
better interactions between polymer chains and clay particles, atic investigation on the role of cations and clay particle size
the added water must be accounted for when measuring the on MFT flocculation is crucial to select appropriate polymers
effectiveness of the flocculant. The net water release (NWR) is and to optimize their dosages to treat tailings efficiently and
calculated with the expression economically. MFT flocculation is a complex process, which
requires the use of statistical tools to allow more systematic
V f − VP studies. Motta et al. used surface response methodologies to
NWR = (3) investigate the impact of cation concentration (Na+ and Ca2+)
VMFT
and anionic PAM dosage, as well as their interactions, in MFT
where Vf is the volume collected, Vp is the volume of water dewatering MFT.[18] They also examined the effect of clay par-
added to the system, and VMFT is the volume of water contained ticle size during the flocculation of samples with distinct par-
in MFT. ticle size distributions (PSDs). Their results are summarized in

Figure 3.  Schematic of FBRM. Reproduced with permission.[13] Copyright 2006, IWA Publishing.

Macromol. Rapid Commun. 2019, 40, 1800644 1800644  (4 of 18) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mrc-journal.de

Table 1.  Summary of chemical structures of polymers. Table 1. Continued.

Polymer Chemical structure Polymer Chemical structure

Anionic polyacrylamide Chitosan-g-CTA

Poly(acrylamide-DADMAC)
Chitosan-g-PAM

P(NIPAM-MATMAC-BAAM)

Bis(3-trimethoxysilylpropyl)
amine (aminosilane)

P(NIPAM-AA-NTBM)

Figure  4. The first row in Figure 4 shows the effect of cation


concentration (Na+ and Ca2+) and anionic PAM dosage on CST.
The dewaterability increased (low CST) with increasing anionic
Hyperbranched- PAM dosage and Na+ concentration, but it decreased (high
functionalized CST) when the concentration of Ca2+ increased, indicating that
polyethylenes (HBfPE) the addition of Ca2+ to MFT promotes water retention in the
Poly(PCL3ChMA) sediments. The second row in Figure 4 illustrates the effect of
cation concentration (Na+ and Ca2+) and anionic PAM dosage
on solids content. Optimum values for solids content were
obtained when Na+ concentration varied from 400 to 800 ppm,
Poly(PLA4ChMA) the Ca2+ concentration was lower than 800 ppm, and anionic
PAM dosage ranged from 1200 to 2000 ppm.
From a fundamental perspective, the effects of cation con-
centration (Na+ and Ca2+) must be isolated from that of anionic
PAM dosage to differentiate the role of each parameter on floc-
culation. Motta and coworkers combined particle size and zeta
potential measurements to investigate the mechanism of floc for-
Poly(PLA4ChA) mation.[18] Figure 5 shows a schematic representation of changes
in particle surface area, zeta potential, and particle size. Increase
in cationicity changed zeta potentials from negative to positive
values, whereas addition of anionic PAM reduced the zeta poten-
tial. Addition of cations did not significantly change particle sizes,
but the addition of anionic PAM increased it considerably. These
PAM-g-PEOMA
observations revealed that cations alone can neutralize the clay
particles, but cannot settle them due to their inability to form
flocs. The synergistic effect of cations and anionic PAM, on the
other hand, can help floc formation and subsequent settling.
The linear copolymer of acrylamide (AM) and diallyl dime-
thyl ammonium chloride (DADMAC) has been explored as a
potential alternative to conventionally used hydrolyzed PAM
PAM-g-PPO (HPAM) to avoid the use of salts.[19] Table 1 shows the chemical
structure of P(AM-DADMAC). The use of P(AM-co-DADMAC)
for MFT treatment and its effective dewatering performance
has been recently patented.[20] However, the effect of polymer
microstructure on the dewatering performance of P(AM-
DADMAC) is still not well quantified. In general, the number
of studies evaluating the influence of copolymer microstruc-
ture on flocculation and dewatering of MFTs is quite limited,
(Continued) although a few studies have been published for other similar

Macromol. Rapid Commun. 2019, 40, 1800644 1800644  (5 of 18) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mrc-journal.de

Figure 4.  Effect of dosage of anionic PAM and concentrations of Na+ and Ca2+ on A) capillary suction time and B) solids content of MFT. Reproduced
with permission.[18] Copyright 2018, Elsevier.

systems, such as pulp wastewater, sludge dewatering, and concentration of the initiator added to the polymerization
model clays.[21–23] Vajihinejad et al. systematically evaluated reactor. The classic free radical polymerization relationship
the influence of copolymer microstructure on flocculation and between the number-average chain length of the polymer and
dewatering behavior of MFT.[19] They modeled the copolymer the initial initiator concentration at low conversions (Equa-
composition of P(AM-DADMAC) using the Mayo and Lewis tion (5)) can be used to control the molecular weight, assuming
equation.[24] chain termination by disproportionation only.[25]

r1 f 12 + f 1 f 2 kt0.5
(2 f kI [I ]0)
0.5
F1 = (4)
r1 f 1 + 2 f 1 f 2 + r2 f 22
2
1 kp km
= + tr (5)
rn [M ]0 (1 − x n ) kp
where F1 is an instantaneous molar fraction of monomer 1
in the copolymer, f1 and f2 are the instantaneous comonomer
molar fractions of monomers 1 and 2 in the reactor, and r1 and where, r n̅ is the instantaneous number-average chain length of
r2 are the reactivity ratios. Equation (4) relates the molar frac- the polymer, kI, kp, kt, and ktrm are rate constants for initiator
tion of monomer 1 in the copolymer to the molar fraction of decomposition, propagation, termination, and transfer to mon-
monomer 1 in the reaction medium. The average molecular omer, respectively, and f is the fractional initiator efficiency.
weight of the copolymer was controlled by varying the initial Further, they investigated the effect of copolymer composition,
molecular weight, and the dosage on the
ISR of sediments and the turbidity of the
recovered water, as shown in Figure 6.
Although flocculants containing posi-
tive charges have shown promising
flocculation performance, additional func-
tionalities can be added to the copolymers
to enhance the dewaterability of their
sediments. Gumfekar and Soares syn-
thesized a terpolymer, poly(N-isopropyl
acrylamide/2-(methacryloyloxy) ethyl
trimethyl ammonium chloride/N-tert-
butylacrylamide) [P(NIPAM-MATMAC-
BAAM)], containing the hydrophobic
BAAM monomer.[26] The chemical struc-
ture of P(NIPAM-MATMAC-BAAM) is
shown in Table 1. They demonstrated
that the hydrophobic BAAM comon-
omer helped in expelling water from the
sediments, while the cationic MATMAC
Figure 5.  Effect of cations and anionic PAM on A) physical state of clays in MFT, B) zeta potential, comonomer promoted charge neutrali-
and C) particle size. Reproduced with permission.[18] Copyright 2018, Elsevier. zation of negatively charged particles

Macromol. Rapid Commun. 2019, 40, 1800644 1800644  (6 of 18) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mrc-journal.de

Figure 6.  Effect of copolymer composition, molecular weight, and dosage of P(AM-co-DADMAC) on A) initial settling rate of MFT and B) turbidity of
recovered water. Reproduced with permission.[19] Copyright 2017, American Chemical Society.

suspended in MFT. Since the fraction of each comonomer in reactivity ratio r.[27] Figure  7 compares the copolymer com-
the terpolymer determined the performance of the functional position measured experimentally and estimated using the
flocculant, it was important to determine and systematically Mayo–Lewis model as a function of feed composition in binary
control the composition of the terpolymer. Gumfekar and copolymers. The investigation showed how polymer reaction
Soares first experimentally measured and later estimated the engineering tools can be used to develop a framework for the
binary reactivity ratios using the Mayo–Lewis equation (Equa- systematic design of multifunctional flocculants.
tion (4)). Thus, they could minimize compositional drift and Use of a hydrophobic comonomer adds a degree of hydro-
guarantee the synthesis of terpolymers with narrow chemical phobicity to the flocculant that can be controlled at the syn-
composition distributions, suitable as model flocculants for thesis stage only. On the other hand, hydrophobicity-on-demand
MFT dewatering. Moreover, they estimated the composition of can have additional benefits in the treatment of tailings. Such
the terpolymers based on the binary reactivity ratios estimated hydrophobicity can be induced in the polymer upon applica-
using Equation (4), using the Alfrey–Goldfinger equation tion of external stimuli such as changes in temperature and
(Equation (6)) pH. Polymers that change conformation from hydrophilic
to hydrophobic above a certain temperature are called lower
FN : FM : FB = f N λN : f M λM : f B λB (6) critical solution temperature (LCST) polymers. Gumfekar
and Soares synthesized a terpolymer flocculant poly(N-
where F is the monomer fraction in terpolymer, f is the isopropyl acrylamide/acrylic acid/N-tert-butylacrylamide)
monomer fraction in the feed, and λ is the function of f and [P(NIPAM-AA-NTBM)], which contained the permanently

Figure 7.  Comparison of copolymer composition measured experimentally and estimated using the Mayo–Lewis model as a function of feed composi-
tion. Reproduced with permission.[26] Copyright 2018, Elsevier.

Macromol. Rapid Commun. 2019, 40, 1800644 1800644  (7 of 18) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mrc-journal.de

hydrophobicity of the benzoboroxole moieties, resulting in


strong hydrophobic interaction above the LCST. The fastest set-
tling rate was obtained at a polymer dosage of 25 ppm, pH 9,
and temperature of 50 °C. The enhanced settling was attributed
to the strong adhesion of benzoboroxole groups to the kaolin
hydroxyl groups and an increase in the hydrophobic force.
Considering the distinct roles of anionic and cationic poly-
mers, Lu et al. flocculated oil sands tailings in two steps.[31]
First, they added commercial A-PAM, Magnafloc-1011, followed
by the addition of the natural polymer chitosan (dissolved in
acidic medium). Their results indicated that the anionic APAM
induced preliminary bridging among the clay particles, but
could not collect a significant fraction of the fine particles.
The subsequent addition of the cationic chitosan neutralized
the fine particles and also increased the floc size, possibly by
inducing the interfloc aggregation. One must note that chitosan
becomes cationic only in the acidic medium; therefore, it needs
to be dissolved in acids such as acetic acid. The addition of
acidic media to oil sands tailings may possibly limit the applica-
Figure 8.  Lower critical solution temperature (LCST) of the homopoly­mer, tion of such new flocculant systems.
copolymer, and terpolymer of the NIPAM. Reproduced with permis-
sion.[28] Copyright 2018, Elsevier.
Linear (homo-, co-, ter-) polymers have performed well for
the dewatering of MFT due to their abilities to form very high
molecular weights and selection of functional comonomers
hydrophobic NTBM comonomer and the thermosensitive that enhance the flocculation. It should also be noted that linear
NIPAM comonomer that becomes hydrophobic above its polymers do have pendant groups that are considered small
LCST.[28] The chemical structure of P(NIPAM-AA-NTBM) is branches due to the intrinsic structure of repeating units. If the
shown in Table 1. P(NIPAM-AA-NTBM) is soluble in water at branches can be controlled by controlled polymerization tech-
room temperature, but it becomes hydrophobic (water insol- niques, these modified flocculants may offer distinct benefits
uble) above 30 °C. The authors added this thermosensitive that are not offered by linear polymers. Some of the benefits
terpolymer to model tailings at room temperature and raised are low viscosity compared to that of similar molecular weight
the temperature to 50 °C to make the flocculant hydrophobic. linear polymers and very high charge density at low molecular
The additional hydrophobicity due to the NIPAM comon- weights.
omer promoted the dewatering of sediments as evident by
CST measurements. The comonomers NIPAM (above LCST)
and NTBM helped expel water out of sediments due to their 4. Branched Polymer Flocculants
hydrophobicity. Although this was an attractive solution, the
copolymerization of NIPAM with other monomers affected Branched flocculants have been synthesized mainly using
the LCST of the flocculant due to changes in the hydrogen multifunctional initiators or controlled polymerization tech-
bonding ability of the polymer chains. P(NIPAM) showed an niques such as atom transfer radical polymerization (ATRP)
LCST of 33 °C, whereas P(NIPAM-AA) and P(NIPAM-AA- and reversible addition–fragmentation chain-transfer (RAFT)
NTBM) showed an LCST of 36 °C and 30 °C, respectively, as polymerization, but other techniques such as catalyzed polym-
shown in Figure 8. erization and macromonomer polymerization have been less
A temperature-responsive copolymer of NIPAM and explored.
2-aminoethyl methacrylamide hydrochloride (AEMA) was Botha et al. made a series of novel hyperbranched func-
used to flocculate and dewater MFT.[29] The linear copolymer tional polyethylene (HBfPE) flocculants consisting of a hydro-
poly(NIPAM-AEMA) was synthesized using free radical polym- phobic backbone containing anionic (carboxylic acid) func-
erization with several copolymer compositions. The copolymer, tional groups ranging from 35 to 50 mol% using a specifically
mixed with MFT at 25 °C, was able to bridge the fine particles designed palladium catalyst.[32] The catalyst enabled the copo-
together, and copolymer adsorption increased further when the lymerization of ethylene and methyl acrylate (MA) to form
temperature was raised to 50 °C. Although the authors claimed the hyperbranched microstructure. Since the MA-function-
an increase in adsorption due to hydrophobic interactions, the alized polyethylenes were hydrophobic, base hydrolysis was
flocculation tests did not show a significant difference in per- performed to generate anions and increase their solubility in
formance between 25 and 50 °C. It was also necessary to add a water. The chemical structure of HBfPE is shown in Table 1.
cationic copolymer, poly(acrylamide-diallyldimethylammonium Their polymers had significantly lower molecular weights and
chloride), to increase the clarity of the supernatant. viscosities than typical PAM flocculants, and were easier to be
A hydrophobic monomer, 5-methacrylamido-1,2-ben- adequately dispersed in high solids MFT. The HBfPE could also
zoboroxole (MAAmBo), was copolymerized with NIPAM be dosed over a wider optimum range and were not sensitive to
and AEMA to increase the hydrophobicity of poly(NIPAM- overdosing, which improves the practical application of these
AEMA).[30] The LCST of the terpolymer decreased due to the polymers, since overdosing often leads to deleterious effects

Macromol. Rapid Commun. 2019, 40, 1800644 1800644  (8 of 18) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mrc-journal.de

Figure 9.  Comparison of changes in floc size after the addition of conventional polyacrylamide and hyperbranched-functionalized polyethylene. Repro-
duced with permission.[32] Copyright 2017, Elsevier.

in flocculation. FBRM comparative studies showed that PAM However, polyester hydrolysis occurred only after keeping the
formed larger flocs, but their size changed over time; HBfPE, tailings sediments at 85 °C for 9 days. Thus, this technology
on the other hand, formed relatively smaller but uniformly still requires further improvements before it can be adopted at
sized flocs, as shown in Figure 9. a commercial scale because heating a vast amount of tailings
One of the strategies to make hyperbranched polymers is to for a longer period is not practically feasible.
use a chain transfer agent (CTA) during the polymerization. Since poly(PCL3ChMA) contained an average of three poly-
Wang et al. prepared a temperature-responsive hyperbranched caprolactone units in the methacrylic ester side chain, degra-
copolymer of AEMA and NIPAM using 4-cyanopentanoic acid dation and its subsequent effect (more water release) on floc-
dithiobenzoate as CTA.[33] The hyperbranched copolymer pro- culation required relatively higher temperatures and longer
moted secondary consolidation at temperatures above its LCST, times. Gumfekar et al. synthesized a similar flocculant with
as compared to similar linear copolymers. They hypothesized an average of two polycaprolactone units in the methacrylic
that hyperbranched structure exposed the buried positive ester side chain, poly(PCL2ChMA), to facilitate degradation at
charges during the polymer coil-to-globule transition, providing lower temperatures and shorter times.[35] To validate the phe-
additional electrostatic interactions and enhancing the capture nomenon of hydrolytic degradation of polyester side chains
of fine particles. in poly(PCL2ChMA), they compared the flocculation per-
In another approach to design branched polymers, mac- formance with a structurally similar, yet hydrolytically non-
romonomers consisting of desired functionalities (length, degradable poly(trimethylaminoethyl methacrylate chloride)
functional groups, charge density) can be synthesized as a [poly(TMAEMC)]. Since the degradable polycaprolactone
building block, which can be further polymerized using a suit- units were reduced to two, the poly(PCL2ChMA) degraded in
able technique to control the molecular weight of the whole 7 days instead of 9 days for poly(PCL3ChMA). This observa-
polymer. In a series of works published in collaboration by the tion confirmed that both the flocculant’s hydrophobicity and
Soares and Hutchinson groups, a short-chain polyester mac- degradation time could be tuned by changing the average
romonomer [polycaprolactone choline iodide ester methacrylate number of polyester units defined in the macromonomer
(PCL3ChMA)] was synthesized and further polymerized to synthesis. The comparison of CST obtained using degradable
form a novel comb-like, cationic flocculant for MFT floccula- poly(PCL2ChMA) and non-degradable poly(TMAEMC) showed
tion.[34] The structure of poly(PCL3ChMA) is shown in Table 1. that the CST of the sediments dramatically decreased after the
Interestingly, the flocculant [poly(PCL3ChMA)] was hydro- degradation of poly(PCL2ChMA) in 7 days (Figure  11). How-
philic in its synthesized state, but became more hydrophobic ever, there was no change in CST of the sediments obtained by
in response to hydrolytic degradation in MFT sediments. The poly(TMAEMC) even after 7 days because the polymer did not
cationic chain ends of poly(PCL3ChMA) accelerated the settling degrade hydrolytically.
rate of oil sands tailings, while partial hydrolysis of the poly- Although reducing the number of degradable polycaprol-
ester grafts exposed the hydrophobic segments that reduced actone units to two reduced the degradation time, the overall
CST by 30%. The schematic structure of macromonomer and degradation conditions (85 °C and 7 days) to achieve better
subsequent polymerization using aqueous free radical polym- water release is still not adequate for field applications. To over-
erization is shown in Figure 10A. Figure 10B shows the mecha- come this difficulty, Younes and coworkers synthesized similar
nism of ex situ hydrolytic degradation of poly(PCL3ChMA) and flocculants replacing caprolactone (CL) with easily degradable
corresponding photos showing solubility effects. This tech- lactic acid (LA) and using acrylate (less hydrophobic) func-
nology combined the material properties of polyesters with tional groups instead of methacrylate (more hydrophobic).[36]
the productivity of free radical polymerization to form dual The structure of their polymers [poly(PLA4ChMA) and
functional flocculants with characteristics that can be easily poly(PLA4ChA)] are shown in Table 1. The sediments obtained
tuned to control flocculation performance, such as polymeric using LA-based flocculants displayed a higher CST (22 s) than
cation density, hydrophobic content, and polymer architecture. that obtained using CL-based flocculants (14 s), since LA is

Macromol. Rapid Commun. 2019, 40, 1800644 1800644  (9 of 18) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mrc-journal.de

Figure 10.  A) Schematic structure of macromonomer (in box) and its polymerization using aqueous free radical polymerization to form comb-like,
cationic polymer, and B) mechanism of ex situ hydrolytic degradation of poly(PCL3ChMA) and corresponding photos showing solubility effects.
Reproduced with permission.[34] Copyright 2016, Wiley-VCH.

more hydrophilic than CL. Poly(PLA4ChMA) degraded at 50 °C the flocculants. Although these branched polymers show some
(in 5 days) as compared to degradation of poly(PCL2ChMA) at promise as high-tech MFT flocculants, scale-up challenges and
85  °C in 7 days. Alternatively, poly(PLA4ChMA) also degraded economic viability still needs to be carefully assessed.
at room temperature in 12 weeks, not requiring the use of
higher temperatures. It is evident from Figure 12 that changing
the type of polyester can have a pronounced effect on the degra- 5. Graft Polymer Flocculants
dation of the flocculant.
It is evident from these studies that the effective use of Unlike branched polymers, graft polymers do not essentially
branched polymers for MFT dewatering requires precise control require the synthesis of building blocks to form a multifunc-
of branch lengths and adequate choice of functional groups in tional flocculant. Often, the purpose of graft polymers, in the

Figure 11.  Comparison of capillary suction time obtained using degradable poly(PCL2ChMA) and non-degradable poly(TMAEMC), before and after the
degradation. Reproduced with permission.[35] Copyright 2017, American Chemical Society.

Macromol. Rapid Commun. 2019, 40, 1800644 1800644  (10 of 18) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mrc-journal.de

Figure 12.  Visual record of MFT dewatering using poly(TMAEMC), poly(PCL2ChMA), and poly(PLA4ChMA). Reproduced with permission.[37] Copyright
2018, American Chemical Society.

context of flocculants, is to have two functionalities (backbone copolymerized AM with already-synthesized PEOMA grafts
and graft) in the same polymer and to modify the rheology using free radical polymerization. A central composite design
in comparison with linear polymers. Typically, grafting from was used to evaluate the effects of comonomer compositions,
and grafting onto approaches are used to synthesize graft poly- graft length (PEOMA), and molecular weights. Their results
mers.[8,38] Additionally, the monomer that forms the backbone revealed that adding hydrophobic PEOMA grafts (>30 wt%)
can be copolymerized in the presence of the already-synthesized to PAM backbones can effectively increase ISR and reduce
grafts to prepare a graft copolymer. CST, but may not capture all of the fine particles, leading to a
Hripko et al. synthesized polyacrylamide-g-poly(ethylene higher supernatant turbidity. This even contradicted the most
oxide methyl ether methacrylate) (PAM-g-PEOMA) with dif- common assumption about molecular weight, that is the higher
ferent comonomer compositions, hydrophobic graft chain (≈106 Da) the molecular weight, the better the flocculation.
lengths, and molecular weights.[39] The chemical struc- PAM-g-PEOMA of low molecular weight (<2 × 105 Da) was still
ture of PAM-g-PEOMA is shown in Table 1. The authors effective at flocculating MFT at optimum conditions. Figure 13

Figure 13.  Effect of polymerization conditions on the CST of PAM-g-PEOMA: A) Ln(CST) versus Log[I] and PEOMA wt% and B) Ln(CST) versus PEOMA
length and wt%. Reproduced under the terms of the Creative Commons Attribution License CC-BY.[39] Copyright 2018, The Authors. Published by
Wiley-VCH.

Macromol. Rapid Commun. 2019, 40, 1800644 1800644  (11 of 18) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mrc-journal.de

Figure 14.  Comparison of the CST of sediments obtained using several PAM-g-PPO samples and commercial PAM at A) 2 wt% and B) 20 wt% MFT.
Reproduced with permission.[40] Copyright 2016, Elsevier.

shows the effect of polymerization conditions on the CST of The technique of graft polymerization is also effective in
PAM-g-PEOMA. Lowering the initiator concentration (higher the case of natural polymers. Many natural polymers such as
polymer molecular weight) and maintaining the PEOMA chitosan, cellulose, alginate, and starch are used as backbones
length between 15 and 22 significantly reduced the CST. The to form graft copolymers useful for flocculation.[8,41,42] Hydro-
results also indicate that the length and percentage of hydro- phobic, anionic, or cationic functional groups can be grafted
phobic PEOMA have to carefully optimized to obtain the best onto natural polymers to transform them in multifunctional
dewatering performance. flocculants.
Further fine-tuning of the properties of graft copolymers Oliveira et al. grafted the small cationic molecule 3-chloro-
may require precise designing of grafts by building macromon- 2-hydroxypropyltrimethylammonium chloride (CTA) and PAM
omers with desired functionalities. Reis et al. synthesized pol- onto chitosan and used them to flocculate MFT.[43] The chem-
yacrylamide-g-poly(propylene oxide) (PAM-g-PPO) by first syn- ical structures of chito-g-CTA and chito-g-PAM are shown in
thesizing hydrophobic PPO grafts with vinyl groups, followed Table 1. Increasing the CTA concentration increased the zeta
by aqueous free radical initiated micellar copolymerization of potential of the polymer, which was beneficial to neutralize
AM and PPO.[40] Micellar polymerization was essential because clays in MFT. Chito-g-PAM settled MFT solids faster than
even though AM was soluble in water, PPO is not. Since the chito-g-CTA because the longer PAM grafts formed larger flocs
amount of hydrophobicity plays a crucial role in MFT floccu- (Figure  16A). The authors also compared the SRF of MFT
lation, they selected PPO with two molecular weights, 300 g
mol−1 and 1000 g mol−1. The chemical structure of PAM-g-PPO
is shown in Table 1. Their PAM-g-PPO flocculants had two
distinct features. First, PAM-g-PPO could be used for MFT as
high solids loadings (20 wt%) with the minor compromise of
dewaterability as compared to diluted MFT (2 wt%). From an
industrial perspective, this is an important feature because sev-
eral polymers cannot dewater MFT efficiently at higher solids
concentrations.
Figure  14 shows the CST of sediments obtained using sev-
eral PAM-g-PPO and commercial PAM at 2 wt% and 20 wt%
MFT. Since flocs formed by PAM are known to hold signifi-
cant water, often filtration or centrifugation are employed in
the industry to obtain stackable solids. However, mechanical
treatments have challenges to remove bound water from PAM-
treated MFT. To evaluate the applicability of PAM-g-PPO in
mechanical treatments, Reis et al. flocculated MFT using PAM-
g-PPO and centrifuged the sediments for 2 min to remove the
water from flocs.[40] Interestingly, they obtained higher solids
percentage (more water removal) using PAM-g-PPO than the
sediments treated with commercial anionic PAM (Figure  15).
The results indicated that PAM-g-PPO graft copolymers could Figure 15.  Solids content of the sediments obtained by flocculation and
significantly improve the densification of MFT when integrated centrifugation using PAM-g-PPO and commercial anionic PAM. Repro-
with mechanical methods such as centrifugation. duced with permission.[40] Copyright 2016, Elsevier.

Macromol. Rapid Commun. 2019, 40, 1800644 1800644  (12 of 18) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mrc-journal.de

Figure 16.  A) Initial settling rate of MFT using chito-g-CTA and chito-g-PAM, and B) specific resistance to filtration of MFT using chito-g-CTA, chito-
g-PAM, and commercial cationic PAM. Reproduced with permission.[43] Copyright 2018, American Chemical Society.

using chito-g-CTA, chito-g-PAM, and commercial cationic PAM for thermal insulation application, compression, tension, and
(C-PAM) (Figure 16B). The graft polymers had the lowest SRF thermal properties were measured. The increase in MFT solids
above 7000 ppm, which showed the tendency of sediments in composite increased the mechanical properties up to a cer-
formed by graft polymers to expel water as compared to C-PAM. tain extent, beyond which they deteriorated. The authors sug-
Several researchers have used ultraviolet (UV) radiation to gested a positive effect due to clay reinforcement, and a nega-
initiate grafting reactions onto the backbones of natural poly- tive effect due to the disruption of hydrogen bonds between
mers.[44–46] UV radiation presumably generates radicals that polymer chains. Further, they also attributed the enhanced dis-
initiate the free radical polymerization of the desired monomer persion of MFT solids to van der Waal and ionic interactions
onto the backbone of another polymer. Recently, Liu and cow- between residual bitumen of MFT and polyol matrix. Figure 17
orkers grafted acrylamide and itaconic acid onto chitosan back- shows the TEM image of PU-MFT composite and schematic
bones using UV and 2,2′-azobis(2-methylpropionamide)dihy- model demonstrating how polymer chains and MFT particles
drochloride initiators.[47] It was unclear why the authors used a interact.
free radical initiator in combination with UV radiation, which The presence of water in MFT substantially hinders most
should generate abundant free radicals on itself. The graft of the polymerizations and thus requires the removal of water
poly­mer was used to flocculate kaolin suspensions with the aid prior to the polymerization of monomers. The removal of water
of Ca2+ ions. A loading of 30 mg L−1 graft polymer was neces- from MFT to use MFT solids in the polymer may defeat the
sary to flocculate a kaolin suspension of 500 mg L−1. purpose of using whole MFT as it is. Moran and Soares formed
packaging films made up of plasticized starch and MFT,
without altering the physical properties of MFT.[49] They plas-
6. Other Tailings Remediation Strategies ticized the starch using glycerol as a plasticizer, a reaction that
requires water as a medium. The addition of MFT to the starch
Most of the literature on MFT remediation focus on the use plasticization facilitated the reaction due to the presence of
of polymers to induce solid–liquid separation. The efficiency water in MFT and did not negatively affect the reaction. More-
of this approach depends on several parameters related to the over, the exfoliation of MFT clays in the starch matrix enhanced
polymer, MFT, and operational conditions. Nevertheless, inef- the mechanical and thermal properties of the composites. They
ficient flocculation may result in the generation of secondary compared the MFT-starch composite with typically used mont-
sludge that requires further action, such as mechanical treat- morillonite-starch composites. MFT-starch composites made
ment. Despite advancements in innovative polymeric floccu- with up to 20% filler content reached the same tensile mod-
lants, the issue of MFT dewatering does not seem to have a ulus achieved by composites made with 5% montmorillonite,
single step solution. which is a positive feature of these composites in terms of MFT
Inspired by polymeric composites wherein filler material utilization and tailing pond land reclamation. The dynamic
improves the properties of the composite, Vajihinejad and mechanical analysis of composites identified a glass–rubber
Soares exploited a novel approach of using MFT as a filler transition, but also the heterogeneity of the composites by the
material to form a composite material with polyurethane (PU) presence of a second transition at higher temperatures. None-
matrix.[48] Previously, PU composites have found applications theless, the use of MFT as a filler material for starch can be a
in cushioning, thermal insulation, footwear, buoyancy prod- cost-effective approach compared to the use of purified clays.
ucts, and packaging. The authors prepared PU-MFT composite Thompson et al. adopted a unique approach of using alkox-
foams by catalyzed polyaddition reactions between polyol and ysilanes to induce the flocculation in MFT.[50] They hypoth-
isocyanate, in the presence of MFT solids and using water as a esized that alkoxysilanes can form oligomers or polymers after
blowing agent to form the foam structure. To use the composite sequential hydrolysis and condensation reactions in aqueous

Macromol. Rapid Commun. 2019, 40, 1800644 1800644  (13 of 18) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mrc-journal.de

Figure 17.  A) TEM image of PU-MFT composite and B) schematic model demonstrating how polymer chains interact with MFT particles. Reproduced
with permission.[48] Copyright 2016, Wiley-VCH.

alcoholic solution. Moreover, they can also bond to other mate- MFT by increasing the settling velocity, hence reducing the size
rials, including those with SiOSi bonds, such as clay min- of the treatment unit while maintaining the high throughput.
erals in MFT via silicate bonding. Alkoxysilanes could polym- Several parameters such as mixing energy, flocculant dosage,
erize and bond with suspended clay particles in MFT, in a initial solids concentration, suspension chemistry, polymer’s
manner similar to conventional polymer flocculants, although physical properties, and geometry of treatment unit (e.g.,
via a different mechanism. Once bonded together by alkox- stirred vessel vs pipe flow) affect the size of aggregates.
ysilane bridges, the clay particles might undergo flocculation Vajihinejad and Soares used a population balance model
if they could be brought close together by the tendency of the (PBM) to describe the flocculation of MFT with poly(vinylbenzyl
alkoxysilane polymer to minimize its surface area. The detailed trimethylammonium chloride) [P(VBTMAC)].[12] The size
experimentation using central composite rotatable design indi- and distribution of aggregates in a flocculating suspension
cated that among the alkoxysilanes tested only those containing were tracked by PBM. Since flocculation involves the simul-
an amine group showed a clear, dose-dependent improvement taneous aggregation and breakage of aggregates, their PBM
of dewatering properties of undiluted MFT. The chemical struc- involved functions for both aggregate and breakage distribu-
ture of aminosilane is shown in Table 1. The positive charges of tion. Equation (7) is the continuous form of PBM as applied to
the alkoxysilanes containing amine groups might have readily flocculation.[51]
attracted to the negatively charged clay particles in MFT by
electrostatic interactions. As a result of this charge neutraliza- ∂n ( v , t ) 1
v

tion, the electrostatic repulsion between clays can be reduced, = ∫ Q ( v − v ′ , v ′ ) n ( v − v ′ , t ) n ( v ′ , t ) dv ′


∂t 20
allowing them to come closer to each other and settle. Although ∞ ∞

there was a noticeable increase in the solids content of sedi- + ∫ Γ ( v , v ′ )S ( v ′ ) n ( v ′ , t ) dv ′ − ∫ Q ( v , v ′ ) n ( v , t ) n ( v ′ , t ) dv ′ − S ( v ) n ( v , t )


ments after aminosilane treatment of MFT, the dewaterability (7)
v 0

measured in terms of CST was still significantly lower than


PAM-based flocculants. Even though this approach did not pro- where n(v, t) and n(v′, t) are number concentrations of aggre-
vide the ultimate solution to MFT dewatering, the combined gates (number per unit volume) of size (volume) v and v′at time
use of alkoxysilanes with other technologies such as polymeric t, Q(v, v′) is the aggregation rate coefficient (aggregation kernel),
flocculants and mechanical treatments may have the potential S(v) is the fragmentation kernel, and Γ(v, v′) is the breakage
to reach the desired dewatering criteria. distribution function (simply how many v-sized daughter
aggregates are formed when a v′-sized aggregate, v′ > v, is frag-
mented). The first two terms in Equation (7) account for the
7. Aggregate Structure Quantification and formation of aggregates of size v, whereas the two last terms
account for the disappearance of aggregates of size v.
Modeling the Floc Formation in Tailings
The aggregation kernel between two aggregates of a size di
7.1. Flocculation Monitoring Using Population and dj was defined as Equation (8).
Balance Approach
Q i , j = β i , j × α i , j (8)
Innovative polymer flocculants for solid–liquid separation in
oil sands tailings are useful only if the flocculation could be where βi,j and αi,j are collision frequency and capture efficiency
adequately optimized and monitored for various conditions. (the probability of aggregation upon collision), respectively. Col-
It is plausible using mathematical models that describe the lision frequency term included the role of fluid motion (mixing)
flocculation process, which is ultimately linked to the physical and aggregates structure, and capture efficiency included the
properties of the aggregates such as size, density, and fractal colloidal forces and hydrodynamic repulsions in the floccula-
dimension. For example, larger aggregates quickly dewater tion suspension.

Macromol. Rapid Commun. 2019, 40, 1800644 1800644  (14 of 18) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mrc-journal.de

Table 2.  Summary of kinetic terms defined in the population balance model to predict flocculation.[12]

Kinetic term Expression

Aggregation kernel Q i , j = 0.162 G (d i + d j )3 [(α max − α min )e − k d t + α min ] (9)


i −1
d i = d 0,1 2 dF

G is the verage turbulent shear rate, dF is aggregates mass fractal dimension, di is the effective dimeter of a fractal aggregate that equivalently
i −1
contains 2 dF smallest primary particle. Term (α max − α min )e − k d t + α min (=αi,j ) was included in the PBM to address the S-shaped (Figure 18)
response of aggregate size evolution in time, with αmax and αmin representing the maximum and minimum capture efficiency, respectively, and
kd is a decay coefficient.
Breakage kernel[52] S i = s1 G s 2 d i (10)
Constants s1 and s2 represent the strength of aggregates
Average shear rate 1
 ε × ρ sus  2 (11)
G=
 µ sus 

( )
−2
N pN 3D 5 ∅
ε= , ρ sus = ρ s ∅ + ρ w (1− ∅), µ sus = µ w 1− eff
V 0.65
ρsus, ρs, ρw, μsus, μw are the density of suspension, solids, water, and viscosity of suspension[53] and water, respectively. ε is fluid average energy
dissipation rate,[54] Np,N, V are impeller power number, rotation speed, and suspension volume, respectively. ∅eff is aggregates effective
volume fraction[55]

3− d F
 d 
∅ eff = ∅  i,ave 
 d 0,i,ave 

where di,ave is average aggregates size (over the whole mixing time), d0,i,ave is average primary particles size obtained from laser diffraction to be
16.5 µm.

Their systematic approach involved four steps: i) defining i −1

the appropriate kinetic terms for PBM, ii) deciding initial d0,i = 0.27 × 2 3 µm (14)
conditions of PBM, iii) solving PBM, and fitting the model to
experimental data, and iv) predicting flocculation using the Solving the model and fitting it with experimental data
model. (step iii) estimated parameters related to aggregation rate,
The kinetic terms defined for PBM (step i) are summarized breakage rate, and fractal dimension. The model was fit to
in Table 2. the experimental data by minimizing the objective function
Laser diffraction technique was used to obtain the data F in Equation (15) using the particle swarm optimization
needed to define the initial conditions for PBM (step ii). Equa- (PSO).
tion (12) gave an initial number concentration of particles (N0,i)
in MFT.[12] max
F (α max , …, s2 ) = ∑ ( dexp − dmodel ) (15)
2

∅ × v ( d0,i ) t =0
N 0,i =
π 3 (12)
d0,i dexp was obtained by calibrating the FBRM chord length data
6
with laser diffraction spectroscopy. dmodel was calculated using
where ∅ is the total volume fraction of primary particles in Equation (16).[12]
MFT. v(d0,i) is the initial volume fraction of particles with a size
d0,i, which was obtained using Equation (13).[12]

max
N i di 4
dmodel = i =1
(16)

max
v ( di +1LD ) − v ( di −1LD ) N i di 3
(d ) + v (d ) (13)
i =1
v ( d0,i ) = 0,i − di −1 LD
i −1
LD

di +1LD − di −1LD
Their model successfully predicted the effect of polymer
where v(di − 1LD) and v(di + 1LD) are the experimental volume frac- dosage on capture efficiency and mean aggregate size of
tion of primary particles obtained by the laser diffraction PSD particles. Moreover, the model also predicted the mean
appearing immediately before and after d0,i, respectively. aggregate size as a function of shear rate (Figure 18). Indus-
The smallest particle size d0,1 in their MFT was 0.27 µm, trially, these parameters are most relevant and the model
so geometrical discretization with progression factor of 2 was could serve as a soft sensor to predict the performance of
obtained by Equation (14).[12] flocculation.

Macromol. Rapid Commun. 2019, 40, 1800644 1800644  (15 of 18) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mrc-journal.de

Figure 18.  Experimental and predicted mean aggregate size as a function of A) shear rate, B) polymer dosage, and C) the effect of dosage on capture
efficiency of the particles in MFT. Reproduced with permission.[12] Copyright 2018, Elsevier.

7.2. Quantification of Aggregate Microstructure where U is terminal settling velocity of the aggregate of equiv-
alent size (dst); μ, ρl, and g are fluid viscosity, fluid density,
Understanding the impact of polymer microstructure on phys- and the gravity constant, respectively. dst was calculated using
ical characteristics of the aggregates allows for the optimum Equation (18).[57]
design of novel flocculants with tuned properties targeted
for the unique environment of oil sands tailings. One of the  2 
 a  a
important characteristics of aggregates is their density, which  0.8248 + 0.168   + 1.033 × 10 −2   
 b  b
directly affects the sedimentation rate, resistance to shear, and dst =   b (18)
 −3  a 
3
−5  a 
4

dewatering.  −1.264 × 10   + 3.728 × 10   
After Botha et al.[32] showed promising flocculation per-  b b 
formance of hyperbranched polyethylenes (HBfPE) in MFT
flocculation, which is described in Branched Polymers sec- Figure  19A compares free settling velocity as a function of
tion, Costine et al.[56] quantified the structure of aggregates size for MFT flocculated with HBfPE polymers and commercial
formed by HBfPE using a combined application of FBRM and anionic PAM, as measured by FDA. Aggregates formed with
floc density analyzer (FDA). They monitored the evolution of HBfPE polymers generally settled faster than those with A-PAM
aggregate size by FBRM over the full duration of the floccula- of an equivalent size, although the estimated molecular weight
tion process. Simultaneously, undisturbed free settling veloci- of the A-PAM was 100 times more than HBfPE, reflecting the
ties and structures of hundreds of individual aggregates were importance of the right chemistry of the flocculant (balance of
quantified by use of a digital camera, connected to image pro- hydrophobicity/hydrophilicity). Figure 19B also compares the
cessing software of FDA. Further, they calculated the density estimated densities of aggregates formed as a function of aggre-
of an individual free settling aggregate (ρa) using Stokes law gate size, derived from the settling data by Equation (14). Aggre-
(Equation (17). gate densities were generally larger for HBfPEs than for A-PAM
at an equivalent size, up to the size of 150 µm. The results indi-
18 µ U cated that MFT treated with the HBfPE polymers had a higher
ρa = + ρl (17)
g dst2 proportion of inter- rather than intra-aggregate liquor, that is,

Figure 19.  A) free settling velocity as a function of aggregate size and B) estimated densities of aggregates formed as a function of aggregate size using
HBfPE polymers and commercial anionic PAM. Reproduced with permission.[56] Copyright 2018, John Wiley & Sons.

Macromol. Rapid Commun. 2019, 40, 1800644 1800644  (16 of 18) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mrc-journal.de

more liquor between aggregates rather than within the struc- [8] S. Wang, L. Zhang, B. Yan, H. Xu, Q. Liu, H. Zeng, J. Phys. Chem. C
tures, as the author suggested, an important implication to the 2015, 119, 7327.
solids dilution in MFT prior to flocculation. [9] M. Scholz, Ind. Eng. Chem. Res. 2005, 44, 8157.
[10] O. Thomas, M.-F. Thomas, UV-Visible Spectrophotom. Water Waste-
water 2017, 27, 317.
[11] J. Gregory, Adv. Colloid Interface Sci. 2009, 147–148, 109.
8. Summary and Outlook [12] V. Vajihinejad, J. B. P. Soares, Chem. Eng. J. 2018, 346, 447.
[13] P. A. Vanrolleghem, B. D. Clercq, J. D. Clercq, M. Devisscher,
Current flocculation technologies for MFT still need much D. J. Kinnear, I. Nopens, Water Sci. Technol. 2006, 53, 419.
improvement from fundamental and industrial perspectives. [14] H. Lu, L. Xiang, X. Cui, J. Liu, Y. Wang, R. Narain, H. Zeng, Lang-
Limitations of emerging polymer-based technologies occur in muir 2016, 32, 11615.
terms of scale-up, cost, effectiveness over conventional mate- [15] C. Wang, C. Han, Z. Lin, J. Masliyah, Q. Liu, Z. Xu, Energy Fuels
rials, or poor understanding of tailings/polymer interactions 2016, 30, 5223.
and their control. Authors also note a major barrier to obtaining [16] A. Sworska, J. Laskowski, G. Cymerman, Int. J. Miner. Process. 2000,
the desired set of performance matrix for new polymers from 60, 143.
[17] D. R. L. Vedoy, J. B. P. Soares, Can. J. Chem. Eng. 2015, 93, 888.
industry to match with their expectations. Most current tail-
[18] F. L. Motta, R. Gaikwad, L. Botha, J. B. P. Soares, Chemosphere
ings management practices employ mechanical treatments 2018, 208, 263.
after polymer-assisted flocculation to reach the desired levels [19] V. Vajihinejad, R. Guillermo, J. B. P. Soares, Ind. Eng. Chem. Res.
of dewatering. However, the sediments generated by floccula- 2017, 56, 1256.
tion do not necessarily suit the mechanical operations and their [20] C. S. Sikes, T. D. Sikes, M. A. Hochwalt, Aquero Co LLC,
working principles. The microstructure of polymers and thus US9321663B2, 2016.
the properties of sediments can be customized to match the [21] H. Ren, W. Chen, Y. Zheng, Z. Luan, React. Funct. Polym. 2007,
requirements of these mechanical operations. As experimental 67, 601.
research on tailings flocculation is rapidly growing, the mode- [22] Y. Zhou, G. V Franks, Langmuir 2006, 22, 6775.
ling studies of flocculation processes are very sparse. Modeling [23] H. Li, J.-P. O’Shea, G. V Franks, AIChE J. 2009, 55, 2070.
[24] F. R. Mayo, F. M. Lewis, J. Am. Chem. Soc. 1944, 66, 1594.
of MFT flocculation process using new polymers is imperative
[25] S. M. Shawki, A. E. Hamielec, J. Appl. Polym. Sci. 1979, 23, 3341.
to understand the interactions among various components in [26] S. P. Gumfekar, J. B. P. Soares, Chemosphere 2018, 210, 156.
MFT and polymer and to validate them. Further, such models [27] C. Hagiopol, Copolymerization: Toward a Systematic Approach,
come in handy for on-field operations and act as soft sensors. Springer, Boston, MA 1999.
Considering the current generation rate of oil sands tailings in [28] S. P. Gumfekar, J. B. P. Soares, Chemosphere 2018, 194, 422.
addition to existing legacy tailings, there is an urgent need for [29] D. Zhang, T. Thundat, R. Narain, Langmuir 2017, 33, 5900.
advancements in terms of innovative polymers, their control [30] H. Lu, Y. Wang, L. Li, Y. Kotsuchibashi, R. Narain, H. Zeng, ACS
strategies, and performance. Appl. Mater. Interfaces 2015, 7, 27176.
[31] Q. Lu, B. Yan, L. Xie, J. Huang, Y. Liu, H. Zeng, Sci. Total Environ.
2016, 565, 369.
[32] L. Botha, S. Davey, B. Nguyen, A. K. Swarnakar, E. Rivard,
Conflict of Interest J. B. P. Soares, Miner. Eng. 2017, 108, 71.
[33] Y. Wang, Y. Kotsuchibashi, Y. Liu, R. Narain, Langmuir 2014, 30,
The authors declare no conflict of interest. 2360.
[34] T. R. Rooney, S. P. Gumfekar, J. B. P. Soares, R. A. Hutchinson,
Macromol. Mater. Eng. 2016, 301, 1248.
[35] S. P. Gumfekar, T. R. Rooney, R. A. Hutchinson, J. B. P. Soares, ACS
Keywords Appl. Mater. Interfaces 2017, 9, 36290.
dewatering, flocculation, oil sands tailings, water-soluble polymers [36] G. R. Younes, A. R. Proper, T. R. Rooney, R. A. Hutchinson,
S. P. Gumfekar, J. B. P. Soares, Ind. Eng. Chem. Res. 2018, 57, 10809.
Received: August 30, 2018 [37] G. R. Younes, A. R. Proper, T. R. Rooney, R. A. Hutchinson,
Revised: October 4, 2018 S. P. Gumfekar, J. B. P. Soares, Ind. Eng. Chem. Res. 2018, 57, 10809.
Published online: November 12, 2018 [38] C. Gao, H. Möhwald, J. Shen, Polymer 2005, 46, 4088.
[39] R. Hripko, V. Vajihinejad, F. LopesMotta, J. B. P. Soares, Glob. Chal-
lenges 2018, 2, 1700135.
[40] L. G. Reis, R. S. Oliveira, T. N. Palhares, L. S. Spinelli, E. F. Lucas,
[1] J. Masliyah, Z. J. Zhou, Z. Xu, J. Czarnecki, H. Hamza, Can. J. Chem. D. R. L. Vedoy, E. Asare, J. B. P. Soares, Miner. Eng. 2016, 95, 29.
Eng. 2004, 82, 628. [41] S. Mishra, A. Mukul, G. Sen, U. Jha, Int. J. Biol. Macromol. 2011,
[2] C. Wang, D. Harbottle, Q. Liu, Z. Xu, Miner. Eng. 2014, 58, 113. 48, 106.
[3] Oil Sands Tailings Technology Review, Technical report, BGC [42] A. Sand, M. Yadav, D. K. Mishra, K. Behari, Carbohydr. Polym. 2010,
Engineering Inc, 2010, https://doi.org/10.7939/R3Z60C18G. 80, 1147.
[4] L. Botha, J. B. P. Soares, Can. J. Chem. Eng. 2015, 93, 1514. [43] L. Pennetta de Oliveira, S. P. Gumfekar, F. Lopes Motta,
[5] H. A. W. Kaminsky, Ph.D.Thesis, University of Alberta, 2008. J. B. P. Soares, Energy Fuels 2018, 32, 5271.
[6] M. MacKinnon, C. Shaw, Int. J. Surf. Min., Reclam. Environ. 2001, [44] L. Chen, C. Liu, Y. Sun, W. Sun, Y. Xu, H. Zheng, Processes 2018, 6, 54.
15, 235. [45] X. Li, H. Zheng, B. Gao, C. Zhao, Y. Sun, Sep. Purif. Technol. 2017,
[7] T. Kasongo, Z. Zhou, Z. Xu, J. Masliyah, Can. J. Chem. Eng. 2000, 187, 244.
78, 674. [46] Y. Sun, M. Ren, C. Zhu, Y. Xu, H. Zheng, X. Xiao, H. Wu, T. Xia,
Z. You, Ind. Eng. Chem. Res. 2016, 55, 10025.

Macromol. Rapid Commun. 2019, 40, 1800644 1800644  (17 of 18) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.mrc-journal.de

[47] B. Liu, H. Zheng, Y. Wang, X. Chen, C. Zhao, Y. An, X. Tang, Sci. [54] D. C. Hopkins, J. J. Ducoste, J. Colloid Interface Sci. 2003, 264,
Total Environ. 2018, 640–641, 107. 184.
[48] V. Vajihinejad, J. B. P. Soares, Macromol. Mater. Eng. 2016, 301, 383. [55] A. R. Heath, P. A. Bahri, P. D. Fawell, J. B. Farrow, AIChE J. 2006,
[49] D. A. Moran, J. B. P. Soares, Can. J. Chem. Eng. 2017, 95, 1901. 52, 1641.
[50] D. K. Thompson, F. L. Motta, J. B. P. Soares, Miner. Eng. 2017, 111, 90. [56] A. D. Costine, V. Vajihinejad, L. Botha, P. D. Fawell, J. B. P. Soares,
[51] S. Kumar, D. Ramkrishna, Chem. Eng. Sci. 1996, 51, 1311. Can. J. Chem. Eng. 2018, 9999.
[52] J. D. Pandya, L. A. Spielman, J. Colloid Interface Sci. 1982, 90, 517. [57] J. Happel, H. Brenner, in Low Reynolds Number Hydrodynamics with
[53] S. Liu, J. H. Masliyah, in Suspensions: Fundamentals and Applications Special Applications to Particulate Media, Mechanics of Fluids and Trans-
in the Petroleum Industry, Advances in Chemistry, Vol. 251, American port Processes, Vol. 1, Martinus Nijhoff Publishers, The Hague 1981,
Chemical Society, Washington, DC 1996, Ch. 3. https://10.1007/978-94-009-8352-6.

Macromol. Rapid Commun. 2019, 40, 1800644 1800644  (18 of 18) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

You might also like