You are on page 1of 11

Research Article

Cite This: ACS Appl. Mater. Interfaces 2017, 9, 36290-36300 www.acsami.org

Dewatering Oil Sands Tailings with Degradable Polymer Flocculants


Sarang P. Gumfekar,†,‡ Thomas R. Rooney,†,§ Robin A. Hutchinson,*,§ and Joaõ B. P. Soares*,‡

Department of Chemical and Materials Engineering, University of Alberta, Edmonton, Alberta T6G 2V4, Canada
§
Department of Chemical Engineering, Queen’s University, Kingston, Ontario K7L 3N6, Canada
*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via UNIV DE CONCEPCION UDEC on June 13, 2020 at 16:15:06 (UTC).

ABSTRACT: We synthesized hydrolytically degradable cationic polymers by micellar radical polymerization of a short-chain
polyester macromonomer, polycaprolactone choline iodide ester methacrylate (PCL2ChMA) with two polyester units, and used
them to flocculate oil sands mature fine tailings (MFT). We evaluated the flocculation performance of the homopolymer and
copolymers with 30 mol % acrylamide (AM) by measuring initial settling rate (ISR), supernatant turbidity, and capillary suction
time (CST) of the sediments. Flocculants made with trimethylaminoethyl methacrylate chloride (TMAEMC), the monomer
corresponding to PCLnChMA with n = 0, have improved performance over poly(PCL2ChMA) at equivalent loadings due to their
higher charge density per gram of polymer. However, MFT sediments flocculated using the PCL2ChMA-based polymers are
easier to dewater (up to an 85% reduction in CST) after accelerated hydrolytic degradation of the polyester side chains. This
study demonstrates the potential of designing cationic polymers that effectively flocculate oil sands tailings ponds, and also
further dewater the resulting solids through polymer degradation.
KEYWORDS: mature fine tailings, degradable polymer, dewatering, flocculation, micellar radical polymerization

1. INTRODUCTION Several researchers have investigated combinations of chemical


Oil sands deposits in the province of Alberta, Canada, are the and mechanical dewatering techniques, such as flocculation
third largest oil reserve in the world, with deposits estimated at using polymers followed by filtration or centrifugation, as
28.3 billion cubic meters.1,2 Although they are a strategic summarized in a recent article.4 Such energy-intensive
resource for Canada, North America, and the global market, the combinations are useful only if the sediments recovered have
extraction of bitumen from oil sands form waste tailings that little water, and the transportation of these high-solid content
cause technical and environmental problems. Typically, these mixtures to tailing ponds remains a challenge.
tailings contain negatively charged clays (∼30−35 wt %), water Ionic polyelectrolyte copolymers, in which at least one of the
(∼65 wt %), and residual bitumen (3−5 wt %).3 The monomers is anionic and/or cationic, are promising candidates
composition and water chemistry of the mixture makes it for MFT treatment.5−7 Cationic polymers are favored because
harder to dewater the tailings than to recover oil from the the clays suspended in tailings have net negative charges.
reserves. Although coarse particles settle quickly, the fine Flocculation of clays occurs through a combination of bridging
particles remain suspended in water, with no further dewatering and charge neutralization: the (generally high molecular
for many decades. The two main challenges associated with
weight) polymer chains “bridge” the solid particles in
such mature fine tailings (MFT) are (i) to recover water
suspension, while the randomly distributed ionic monomers
trapped in the tailings, and (ii) to further dewater the sediments
obtained after the flocculation. Existing technologies use water- in the polymer chains neutralize the charges on the clay surface.
soluble polymers, such as polyacrylamide (PAM) and its ionic However, the flocs still contain water because the polymer
copolymers, to form large flocs and induce solid−liquid flocculants are hydrophilic. Some researchers are addressing
separation. However, because of the strong water-affinity of this problem by synthesizing hydrophobically modified
PAM, the flocs entrap a significant amount of water, forming
low-shear strength sediments that are difficult to dewater. Received: July 15, 2017
Research in this area aims at improving dewatering Accepted: September 25, 2017
technology and developing more effective flocculating agents. Published: September 25, 2017

© 2017 American Chemical Society 36290 DOI: 10.1021/acsami.7b10302


ACS Appl. Mater. Interfaces 2017, 9, 36290−36300
ACS Applied Materials & Interfaces Research Article

polymers,8−13 because hydrophobic groups present in the Scheme 1. Chemical Structures of Cationic Macromonomers
flocculant cause the sediments to retain less water. Used in Radical Homopolymerizations, As Well As
Functionalization of hydrophobic polymers such as poly- Respective Copolymerizations with Acrylamide (AM) To
ethylene is one of the approaches used to maintain a balance Produce Flocculants Investigated in This Work
between hydrophilicity and hydrophobicity within the poly-
mer.10 Often, such hydrophobically modified polymers are
synthesized by surfactant-assisted emulsion polymerization or
precipitation polymerization.14,15 However, the presence of
surfactants in the flocculants is undesirable as they form froth
during flocculation.
A range of innovative flocculating agents are currently under
investigation, including thermoresponsive, pH-responsive, and
CO2-switchable copolymers synthesized with cationic mono-
mers, as well as hydrophobically modified monomers.6,16−20
Although using polymers that can reversibly change from modified through copolymerization with a common water-
hydrophilic (for efficient flocculation of the MFT) to soluble monomer such as acrylamide (AM).
hydrophobic (for efficient release of the water remaining in In this work, we systematically investigated the influence of
the sediment) is promising, significant challenges remain the copolymer structure on MFT flocculation behavior to
regarding their implementation in the field. For example, follow up the initial promising results demonstrated with
thermoresponsive polymers such as poly(N-isopropylacryla- kaolin,25 the principle clay component in oil sands mixtures.
mide) (PNIPAM) and its ionic derivatives have been studied The average number of degradable units in the polyester side
due to their temperature-sensitive switching from hydrophilic chain was adjusted to 2 to afford a homopolymer with both
to hydrophobic, but the associated energy costs limit their appreciable cationic density and hydrophobic content, while
large-scale implementation.21 pH-Responsive polymers that still maintaining its hydrolytic degradation characteristics. As
switch from hydrophilic to hydrophobic upon the addition of well as moving to the more challenging MFT system, we
acid or alkali are also attractive, but the amount of acid or base investigated the influence of copolymerizing PCL2ChMA with
required to alter pH may limit their practical applications.22 AM to further control polymer charge density and hydro-
Researchers have also explored the use of CO2-switchable phobicity; a recent study determined that 60−70 mol % AM
flocculants such as N, N, N′, N′-tetramethyl-1,4-diaminobutane, content in a cationic polymer provides a good balance of
imidazolyl cellulose nanocrystals, long-chain alkylamidines and flocculation parameters. 5 The main feature of poly-
poly(dimethylamino)ethyl methacrylate (PDMAEMA)23,24 (PCL2ChMA) − its increased hydrophobicity in response to
that become cationic upon exposure to CO2. However, hydrolysis which does not require external triggers such as
considering the 176 km2 area occupied by tailings ponds in changes in temperature or pH − is demonstrated by comparing
Alberta, mass transfer limitations associated with purging CO2 the MFT sediment dewaterability, as characterized by capillary
and maintaining those conditions may not be a feasible solution suction time, immediately following flocculation to that after 1
either. week of accelerated degradation. In addition, the performance
The approach we report herein involves the synthesis of a of the material is compared to that of a structurally similar
flocculant that is cationic and water-soluble in the “as- cationic copolymer produced by radical copolymerization of
synthesized” state, but contains polyester grafts that degrade trimethylaminoethyl methacrylate chloride (TMAEMC) as well
by hydrolysis to increase the flocculant’s hydrophobicity. In a as nonionic AM (see Scheme 1), whose hydrolytic degradation
previous publication we showed that this strategy enhanced the products are not expected to elicit an increase in polymer
dewaterability of “model tailings” sediments consisting of pure hydrophobicity.
kaolin.25 The novel material is produced by radical polymer-
ization of a short-chain polyester macromonomer, polycapro- 2. MATERIALS AND METHODS
lactone choline iodide ester methacrylate (PCLnChMA), where 2.1. Materials. ε-Caprolactone (CL, 97%), 2-(dimethylamino)-
n is the average number of polycaprolactone units in the ethanol (De, > 98%), stannous octoate (Sn(oct)2, 92.5−100.0%)
methacrylic ester side chain. The cationic tips of the resulting triethylamine (TEA, ≥ 99.5%), basic alumina (Brockmann 1), methyl
comb-like polymer accelerate the settling rate of the oil sands iodide (ICH3, 99%), acrylamide (AM, ≥ 99%), 2,2-azobis(2-
tailings, while partial hydrolysis of the polyester grafts expose methylpropionamidine)dihydrochloride (V-50, 97%), [2-
(methacryloyloxy)ethyl] trimethylammonium chloride solution
the hydrophobic segments, increasing the dewaterability of the
(TMAEMC; 80 wt % in H2O), and nonionic polyacrylamide (PAM,
sediments over time. 5−6 million Da) were purchased from Sigma-Aldrich and used as
The structure of the macromonomer, shown in Scheme 1, received. Tetrahydrofuran (THF, > 99%, ACP Chemicals), anhydrous
differs from trimethylaminoethyl methacrylate chloride diethyl ether (≥99.0%, ACP Chemicals), chloroform-d (CDCl3, 99.8%
(TMAEMC) by the degradable polyester spacers in the side D, Cambridge Isotope Laboratories), and deionized water (Millipore
chain. Radical polymerization of the macromonomer yields Synergy water purification system) were used as received.
comb-polymers whose hydrolytic degradability is controlled by Methacryloyl chloride (MACl, 97%, Sigma-Aldrich) was distilled
the number and type of polyester units specified in the immediately before use to remove reactive dimers. Mature fine tailings
macromonomer synthesis, a feature which has been demon- (MFT, 33.5% solids by mass) obtained from Syncrude Canada Ltd.
were diluted to 5 wt % total solids for the flocculation experiments.
strated for similar hydroxyl-functionalized polyester macro- 2.2. Macromonomer Synthesis. The synthesis of polycaprolac-
monomers investigated for biomedical applications.26,27 The tone 2-(N,N-dimethylamino)ethyl ester methacrylate (PCL2DeMA)
number and type of polyester units in the macromonomer with average n = 2 was performed according to a previously published
structure also affects the hydrophobicity and cation density of procedure modified to reduce reaction time.25 A catalyst/monomer
the resulting comb-polymer; such properties can be further mixture with molar ratio of 1:500 consisting of Sn(oct)2 (35.1 mg, 86.6

36291 DOI: 10.1021/acsami.7b10302


ACS Appl. Mater. Interfaces 2017, 9, 36290−36300
ACS Applied Materials & Interfaces Research Article

Figure 1. Schematic representation of the flocculation procedure.

μmol) and CL (4.94 g, 43.3 mmol), was loaded into a 50 mL sealed refractive index, and light scattering. Two columns (TKS gel
round-bottom flask, purged with nitrogen, and then heated to 130 °C. G6000PW XL-CP) connected in series were used to obtain a better
De (1.93 g, 21.7 mmol) was added to the catalyst/monomer mixture resolution and increase the detection range of the instrument. The
by syringe and allowed to react for 110 min at 130 °C to yield system was calibrated using poly(ethylene oxide) standards provided
polycaprolactone 2-(N,N-dimethylamino)ethyl ester (PCL2De) with by Agilent Technologies. Water containing 0.3 M sodium nitrate and
number-average n = 2.0 and CL conversion ≈94% (as determined by 0.01 M monosodium phosphate was used as a mobile phase for the
1
H NMR in Figure S1). analysis.
Next, PCL2De (6.87 g, 21.7 mmol −OH) was dissolved in 35 mL 2.4. Ex Situ Degradation Study. The poly(PCL2ChMA)
tetrahydrofuran (THF) in a sealed 3 neck 100 mL round-bottom flask, homopolymer was diluted to 1 wt % in H2O and held at 85 °C for
to which 24 mL triethylamine (TEA, 173.0 mmol) was then added. 1 day increments then cooled to room temperature for analysis before
The solution was cooled to 0 °C using an ice bath, bubbled with N2 for increasing the temperature back to 85 °C. The solution pH was
10 min, and then 2.8 mL freshly distilled methacryloyl chloride measured using a Mettler Toledo SevenExcellence pH meter, whereas
(MACl, 28.2 mmol) diluted by 4.1 mL of THF was fed over 1 h using dispersion size was determined with a Malvern Zetasizer Nano ZS
a glass syringe. The reaction mixture was maintained at 0 °C for an (size range 0.3 nm −10 μm) at 25 °C with backscattering optics
additional 3 h, filtered to remove the TEA salt, and then passed (173°), using a 4 mW He−Ne (633 nm) laser. All measurements were
through a column of basic alumina. The solvent was evaporated in made in quartz cuvettes, and the reported sizes represent an intensity
vaccuo to afford 6.19 g of PCL2DeMA in 74% yield including MACl average of at least 30 scans.
impurities (Figure S2). 2.5. Oil Sands Flocculation Tests. The flocculant dosage was
PCL2DeMA (6.19 g, 16.1 mmol) was dissolved in 140 mL diethyl varied from 1000 to 5000 ppm (ppm), on a weight basis relative to the
ether, cooled to 0 °C, and kept under constant flow of nitrogen. weight of solids in the MFT. Figure 1 shows the schematic illustrating
Approximately 3 mL of ICH3 was injected by syringe then the reaction the flocculation procedure. For each flocculation test, 100 mL of a 5 wt
was allowed to warm to room temperature and proceed for 48 h. The % solids mixture was prepared by diluting MFT with deionized water.
white waxy precipitate was collected by filtration, washed three times The pH of the diluted MFT varied between 7.9 and 8.2. This
with 200 mL of cold diethyl ether, and dried under vacuum at 40 °C suspension was transferred to a 250 mL baffled beaker (7 cm diameter
overnight to afford 5.48 g of polycaprolactone choline iodide ester and 9.5 cm height) in which the polymer flocculant and the MFT
methacrylate (PCL2ChMA) (77% methylation yield). According to suspension were mixed in two stages using 316 stainless steel 4-blade
the 1H NMR in Figure S3, the PCL2ChMA macromonomer contains propeller with 2 in. diameter: (1) the MFT mixture was stirred at 600
up to 10 mol % [2-(methacryloyloxyl)ethyl]trimethylammonium rpm for 2 min, followed by addition of the desired quantity of the
iodide (TMAEMI), resulting from unreacted De initiator in the polymer flocculant; (2) mixing was maintained at 600 rpm for an
ROP step. additional 5 min, followed by 2 min at 300 rpm. The mixture was
All 1H NMR characterizations (Figures S1−S3) were performed on immediately transferred to a 100 mL graduated cylinder, and the
a Bruker Avance instrument operating at 400 MHz. The oligomeric mudline (solid−liquid interface) was recorded during the initial
distributions of PCL2DeMA (Figure S4) was assessed by size exclusion settling period.
chromatography (SEC). The SEC setup consists of a Waters 2960 Three measurements were used to assess flocculation performance
separation module instrument with a Waters 410 differential of the polymers: (1) initial settling rate (ISR), (2) supernatant
refractometer (DRI), and four Styragel columns (HR 0.5, 1, 3, 4) turbidity, and (3) capillary suction time (CST). ISR was calculated
maintained at 35 °C with distilled THF as eluent at 0.3 mL min−1. The from the slope of an initial linear portion of the settling curves.28,29 An
DRI detector was calibrated using 14 narrow poly(MMA) standards ISR value of zero was assigned when the MFT suspension did not start
(302−853 000 Da). settling upon addition of the polymer solution. After the settling, the
2.3. Macromonomer Copolymerization. For homopolymeriza- supernatant was collected and the turbidity was measured using a
tions, 2.0 g of PCL2ChMA or TMAEMC was dissolved in 18.0 g of Hach 2100AN turbidimeter. Capillary suction time (CST) of the
H2O (containing 0.22 and 0.40 wt % V-50, respectively) and bubbled consolidated solids remaining after removal of the supernatant was
with nitrogen for 1 h in a sealed 50 mL single neck round-bottom measured with a Triton Electronics 319 multipurpose CST apparatus.
flask. Next, the solution was heated to reaction temperature (70 °C for Pipetting a fixed volume of the solids suspension gave more
PCL2ChMA and 50 °C for TMAEMC), allowed to proceed for 2 h reproducible results than scooping an equivalent portion of the
then cooled to room temperature and stored in the refrigerator at 4 sediment with a spatula.
°C. The same conditions were employed for the copolymerizations, The remaining sediments were kept in the graduated cylinder,
except that the 2.0 g of (macro)monomer loading comprised an which was sealed with a rubber cork to prevent water from
amount of AM which corresponds to 30 mol % cationic (macro)- evaporating. The mixture was held at 85 °C for 1 week to allow the
monomer (56 and 76 wt % TMAEMC and PCL2ChMA relative to accelerated in situ degradation of the polymer within the sediments.
AM, respectively). In all cases, quantitative conversions were verified The CST of these samples was measured after 1 week to assess the
by complete disappearance of the vinyl signals by 1H NMR. effect of polymer degradation on the dewaterability of the sediments.
Each synthesized polymer was characterized using gel permeation The morphology of the solids in the sediments obtained after
chromatography (GPC) (1260 Infinity Multi-Detector GPC/SEC flocculation, as well as the raw MFT, were analyzed using cryo-field
System, Agilent Technologies) to analyze its molecular weight and emission-scanning electron microscope (FE-SEM). The samples were
dispersity (Đ). The GPC system included three detectors: viscosity, prepared by lyophilizing the raw MFT and the sediments followed by

36292 DOI: 10.1021/acsami.7b10302


ACS Appl. Mater. Interfaces 2017, 9, 36290−36300
ACS Applied Materials & Interfaces Research Article

gold coating of 8 nm using Denton Gold Sputter Unit. FE-SEM suspension to 2, 5, and 10 wt % solids to determine the
images were obtained using a Zeiss Sigma FESEM w/EDX and EBSD appropriate conditions for carrying out the flocculations.
at accelerating voltage of 10 kV using an inlens detector. Deionized water was used for all dilutions to reduce the effect
of water chemistry on consolidation. A concentration of 5 wt %
3. RESULTS AND DISCUSSION MFT solids was considered good enough to measure settling
3.1. Synthesis of PCL2ChMA and TMAEMC-Based rates: at 2 wt % MFT, settling was too fast to be measured
Polymers. All four cationic copolymers were synthesized by consistently, and at 10 wt % the settling was too slow, especially
batch radical polymerization in aqueous solution. Although at lower flocculant concentrations.
TMAEMC and AM polymerize in solution, the added The Richardson−Zaki batch settling equation (eq 1) for
hydrophobicity of the PCL spacers makes PCL2ChMA a multiparticle systems support these observations4
reactive surfactant that, in an aqueous environment, propagates
by micellar polymerization.25 Although all polymers have Vp
= (1 − αp)n #
similar number-average molecular weights (Mn) (Table 1), Vp ∞ (1)

Table 1. Polymer Charge Densities, Molecular Weight where Vp is the settling velocity of a particle in suspension
Averages, and Dispersities (Effective Values Relative to PEO relative to a stationary observer in m/s, Vp∞ is the terminal
Calibration)a velocity of a particle in a stagnant fluid, αp is the volume
fraction of the particles, and n is an index that depends on
charge density Mn Mw Reynolds number, with n = 4.7 in Stokes’ regime. According to
polymer (mmol/g) (g/mol) (g/mol) Đ
eq 1, as the fraction of particles in the suspension increases, the
poly(PCL2ChMA) 1.9 430 000 770 000 1.79 settling velocity of particles decreases. In Figure 2, the images at
poly(PCL2ChMA- 1.4 522 000 939 600 1.80
AM)
poly(TMAEMC) 4.8 364 000 625 000 1.72
poly(TMAEMC- 2.7 456 000 842 000 1.82
AM)
a
Copolymers contain 70 mol % AM.

the four copolymers differ primarily in their charge densities


which is defined during the macromonomer synthesis step and
is independent of the number of cationic (macro)monomers in
the final homopolymer. With a repeat unit with molecular
weight of 207.7 Da, the charge density of the poly(TMAEMC)
is 4.8 mmol/g, more than double the 1.9 mmol/g of the
poly(PCL2ChMA), which has an average repeat unit with
molecular weight of 527 Da. The cationic charges of
poly(PCL2ChMA), however, are located farther away from
the chain backbone due to the comb-like structure formed by
polymerization of the macromonomer. The charge densities are
reduced to 2.7 mmol/g for poly(TMAEMC-co-AM) and 1.4
mmol/g for poly(PCL2ChMA-co-AM) by addition of 70 mol %
AM as a comonomer to the chains. The addition of AM also
increases the hydrophilicity of the polymer chains. High-
molecular-weight PAM homopolymer was used as a control in Figure 2. Mudline height after flocculation using 5000 ppm of
this study; although of higher molecular weight (5−6 million poly(PCL2ChMA) with MFT diluted to (a) 2 wt % (30 min); (b) 5 wt
% (30 min); (c) 10 wt % (24 h); and (d) control sample with 5 wt %
Da), the PAM control contained no cationic charges.
MFT and 5000 ppm of PAM (30 min).
The PCL2ChMA macromonomer has iodide counterions,
whereas the commercial TMAEMC has chloride counterions.
As the efficiency of “salting-out” for iodide is greater than for 2 and 5 wt % MFT were taken 30 min after polymer addition,
chloride ions,30 the counterion plays a role in determining the whereas the image at 10 wt % MFT was taken after 24 h. These
ease of adsorption of charged polymer onto the surface of the three settling experiments were conducted with 5000 ppm
clay particles. Moreover, Tian et al. demonstrated that larger poly(PCL2ChMA) as flocculant, and are compared to a 5 wt %
monovalent counterions led to a higher cationic density along MFT settled with 5000 ppm nonionic PAM (control) which
the polymer chain (i.e., less counterion condensation) and thus produced a much more turbid supernatant. In contrast, the
to more extended molecular chains that resulted in efficient cationic poly(PCL2ChMA) destabilizes the MFT by adsorbing
adsorption on the clay surface without extending loops and and neutralizing the negative charges on the clay particles,
tails.31 Therefore, the iodide counterion of poly(PCL2ChMA) inducing floc growth and faster settling rate, as well as low
may be an advantage over polymers with chloride counterions. supernatant turbidity. This result is supported by other studies
3.2. Effect of Solids Content in MFT. The as-received that demonstrate that cationic polymers are better flocculants
MFT contained 33.5 wt % solids and a gel-like consistency that for oil sands tailings than neutral PAM.5,33−35
made it difficult to mix with polymer flocculants. Flocculation 3.3. Initial Settling Rate (ISR). Figure 3 shows the ISR
attempts using the raw MFT did not result in settling of solids. values measured at polymer dosages varying from 1000 to 5000
Alamgir et al. have reported similar observations with MFT ppm. The settling rates for experiments conducted with 1000
containing 31 wt % solids.32 Therefore, we diluted the MFT ppm of poly(PCL2ChMA) as well as 1000 and 2000 ppm of
36293 DOI: 10.1021/acsami.7b10302
ACS Appl. Mater. Interfaces 2017, 9, 36290−36300
ACS Applied Materials & Interfaces Research Article

Figure 3. Initial settling rates (ISR) using PCL2ChMA and TMAEMC-based polymers and copolymers at varying dosages with 5 wt % MFT.
Missing data for poly(PCL2ChMA) and poly(PCL2ChMA-AM) denotes insignificant settling. Inset images show the flocculation behavior after 24 h
at respective dosages.

poly(PCL2ChMA-AM) were too low to be measured, likely final mudline heights are presented in Table S1. Although the
because of the lower charge densities of these polymers TMAEMC homo and copolymers flocculated MFT even at
combined with their low dosages. Even though ISR does not 1000 PPM, the sediments are not as compact (higher final
vary with polymer dosage in a clear trend, all polymers mudline height) as those flocculated with PCL2ChMA
flocculated MFT at reasonable rates. Poly(PCL2ChMA) polymers, with differences seen throughout the 1000−5000
initiated settling at 2000 ppm whereas poly(PCL2ChMA-co- ppm concentration range. The more compact sediments
AM) initiated settling at 3000 ppm, likely due to its lower generated by the PCL2ChMA flocculants may be attributed
charge density.36 In addition, the AM-containing copolymers to the hydrophobicity of the polycaprolactone units in the
have increased hydrophilicity, which causes a decrease in floc macromonomer.
density (and thus ISR) as the polymer concentration is 3.4. Supernatant Turbidity. The clarity of the water
increased further. The distributions of floc formation using recovered after the flocculation was quantified by turbidity
2000 ppm of poly(PCL2ChMA) and poly(PCL2ChMA-co-AM) measurements, with a lowering of the turbidity indicating that
are compared in Figure S5, revealing that poly(PCL2ChMA) the polymer was better able to neutralize and bridge the fine
formed larger flocs than poly(PCL2ChMA-co-AM). particles present in MFT. Figure 4 shows that the supernatant
Although poly(PCL2ChMA) and poly(TMAEMC) are turbidity drops when the dosage of PCL2ChMA polymers
structurally similar, the additional hydrophilicity and greater increased. The turbidity for poly(PCL2ChMA-co-AM) is always
charge density of poly(TMAEMC) aids the adsorption of higher than that of the homopolymer at the same polymer
polymer chains to clay particles. Thus, the combination of dosage, a difference that can be attributed to the lower charge
charge neutralization and bridging increases the ISR of density of the copolymer (1.4 versus 1.9 mmol/g). The
poly(TMAEMC) compared to poly(PCL2ChMA), especially turbidity also decreases sharply when the dosage of
at the lower dosages. The inset pictures in Figure 3 were taken PCL2ChMA polymers increases from 3000 to 4000 ppm,
after 24 h, but the settling profiles leveled off in less than 30 with a smaller improvement seen at 5000 ppm.
min. More importantly, the reported ISR values are comparable The influence of the dosage of TMAEMC polymers on
to those reported for other cationic and anionic polymers in the turbidity differs in important aspects from that of PCL2ChMA
literature under similar conditions.10,35,37 polymers. With a significantly higher charge density of 4.8
The mudline heights in the inset photographs of Figure 3 mmol/g, even 1000 ppm of poly(TMAEMC) leads to a
indicate how compact the sediments are. Numerical values of turbidity lower than 10 NTU, dropping below 1 NTU at a
36294 DOI: 10.1021/acsami.7b10302
ACS Appl. Mater. Interfaces 2017, 9, 36290−36300
ACS Applied Materials & Interfaces Research Article

Figure 4. Supernatant turbidity obtained after flocculation of 5 wt % MFT using PCL2ChMA and TMAEMC homopolymers and copolymers at
varying dosages. Turbidity values are indicated for each experiment, as the scale varies for each plot; “broken” y-axes are used to visualize the relative
changes among the lower turbidity values.

dosage of 3000 ppm, before increasing at higher polymer Capillary suction time (CST) was used to measure the
dosages. Poly(TMAEMC-co-AM) follows a similar trend, dewaterability of the settled sediment. While a lower CST value
although its lowered charge density leads to higher turbidity indicates more efficient dewatering of sediments, previous
at 1000 ppm. In addition, the increase in turbidity seen at 4000 studies in our group have shown that the CST of MFT
and 5000 ppm for the TMAEMC copolymer is not as large as suspensions flocculated with PAM increased for higher polymer
for the homopolymer. The increased turbidity observed at 4000 dosages because PAM is hydrophilic.10 Thus, we expected that
and 5000 ppm of the TMAEMC polymers may be attributed to the more hydrophobic PCL2ChMA polymers would make
excessive charge density that causes the restabilization of sediments with CST values lower than those of TMAEMC
particles above the optimum dosage, as has been noted by other polymers. This expectation, however, was not met, as seen in
researchers.38 We did not observe this type of behavior when the “Before Degradation” results in Figure 5: all sediments
applying the PCL2ChMA flocculants within the dosage range flocculated with poly(TMAEMC) and poly(TMAEMC-co-AM)
we investigated, as the cation charge density was tuned to had CST lower than 20 s for dosages at or above 2000 ppm,
neutralize clay particles in MFT, which exhibit a native zeta compared to 20−40 s for the most favorable 4000−5000 ppm
potential of −38 mV.34 Despite having significantly lower levels of poly(PCL2ChMA) and poly(PCL2ChMA-co-AM).
molecular weights than the commercial PAM tested in this Because the sediments produced with PCL2ChMA were already
investigation, the cationic polymers are effective in overcoming more compact (lower mudlines in Figure 3, and thus lower
the electrostatic repulsion between negatively charged clay water content), perhaps this is not surprising as it may seem at
particles, as well as bridging between particles, both necessary first sight. The solids content of the sediments measured before
features of a successful polymeric flocculating agent.39 and after the flocculation showed that PCL2ChMA produced
3.5. Capillary Suction Time and Degradation Study. sediments with higher solid content than those produced with
While the four cationic polymers tested herein differed in ISR TMAEMC (see Table S2). Further, we also calculated the neat
and turbidities (mostly attributable to differences in charge net water release (nNWR), which is defined as the percent of
density), they all performed well with the challenging MFT the water released that was part of the original MFT sample.
system. The most encouraging finding is that compact The results showed that PCL2ChMA-based polymers released
sediments and clear supernatants were obtained at relatively more water than TMAEMC-based polymers. Sawalha and
low polymer dosages5,8 of less than 5000 ppm (0.5 wt %). Scholz have reported similar observations in their systematic
36295 DOI: 10.1021/acsami.7b10302
ACS Appl. Mater. Interfaces 2017, 9, 36290−36300
ACS Applied Materials & Interfaces Research Article

Figure 5. Capillary suction times for sediments obtained from flocculation of 5 wt % MFT using the four cationic (co)polymers, measured before
and after accelerated degradation of the sediments at 85 °C for 1 week.

study of varying solids contents on CST.40 The CST values our previous work for poly(PCL3ChMA). Although a trans-
measured (also summarized in Table S3) are in a similar range parent solution at the beginning, the partially hydrolyzed
as those reported by other researchers for MFTs studied with poly(PCL2ChMA) completely precipitated from solution after
either lower solid content or higher polymer dosages,8,41 7 days at 85 °C, whereas poly(PCL3ChMA) only precipitated
demonstrating that our new polymers are competitive with after 9 days−this observation confirms that both the flocculant’s
previous flocculants. hydrophobicity and degradation time can be tuned by the
For this particular study, the best results are obtained for a average numbers of polyester units defined in the macro-
poly(TMAEMC) dosage of 3000 ppm, which provides fast monomer synthesis. Herein, we examine the impact of the
settling, very low supernatant turbidity, and a low CST value. increased hydrophobicity of poly(PCL2ChMA) (co)polymers
Adding AM as a comonomer provides a way to control the on the dewaterability of the MFT sediments.
charge density (and perhaps cost) of the polymer, but it Thus, in addition to measuring CST values of the sediments
requires a slightly increased polymer dosage (3000−4000 ppm) after 24 h, we repeated these measurements after aging the
to achieve the same performance.
samples for 1 week at 85 °C, as detailed in the Experimental
Higher levels of the poly(PCL2ChMA), 5000 ppm, are
Section. The results are also summarized in Figure 5, providing
needed to achieve similar results. While the sediment mudlines
a direct comparison of the CST of the sediments obtained
are noticeably lower than for 3000 ppm poly(TMAEMC), the
supernatant turbidities and CST values are slightly higher. immediately after the flocculation (before degradation) and
However, an important feature of PCL2ChMA polymers is that after the in situ accelerated degradation. A clear contrast can be
they can degrade. As shown in our previous study, the partial seen between the sediments flocculated with the PCL2ChMA
hydrolysis of the polyester units reveals hydrophobic segments polymers, for which CST decreased remarkably after
that reduced CST of kaolin clay sediments by 30%.25 The degradation, and the TMAEMC polymers, which showed no
partial hydrolytic degradation of poly(PCL2ChMA) to yield a or little change in dewaterability upon aging. In all cases,
more hydrophobic flocculant was qualitatively confirmed irrespectively of polymer dosage or whether the sample was
during an accelerated ex situ degradation study of 1 wt % poly(PCL2ChMA) or poly(PCL2ChMA-co-AM), the CST
flocculant in H2O at 85 °C. As summarized by Figure S6, the decreased to 10−15 s after 1 week of accelerated degradation.
decreasing pH indicates the continuous release of acidic species, More importantly, these results show that increased hydro-
while the increasing dispersion size can be attributed to an phobicity in response to hydrolysis can be achieved without the
aggregation mechanism which was described in more detail by strict requirement of an external trigger such as a temperature
36296 DOI: 10.1021/acsami.7b10302
ACS Appl. Mater. Interfaces 2017, 9, 36290−36300
ACS Applied Materials & Interfaces Research Article

Figure 6. Cryo-FE-SEM images of: (A) raw MFT, (B) sediments obtained after flocculation, (C) internal structure of a floc, and (D) the sediments
obtained after the polymer hydrolysis.

or pH change (although these factors can certainly accelerate typical structure consisting of stacked layers of clay platelets
hydrolytic degradation). bound by strong surface forces (Figure 6A), typical of
Depending on the pH of the medium, the hydrolysis of the aluminosilicate clays.42 The sediments of MFT treated with
polyester side chains may yield degraded polymer with 5000 ppm of poly(PCL2ChMA) showed a much less ordered
carboxylate or carboxylic acid functionalized pendant groups. structure, as observed in Figure 6B. We speculate that
To elicit the increased polymer hydrophobicity in response to poly(PCL2ChMA) synergistically assisted the exfoliation of
hydrolysis documented in our previous work,27 the majority of intercalated clays in untreated MFT, as the cation charge
the pendant groups must be the carboxylic acid functionalized density is sufficient to overcome the interfacial forces among
moieties obtained under acidic conditions.25 Even though MFT the clays. At the same time, destabilized clay particles are
are mildly alkaline (pH 8−9), the extent of hydrolysis after the “bridged” together by polymeric chains, forming flocs of
accelerated degradation test must have been sufficient to approximate dimensions 25 μm × 12 μm, as also observed by
sufficiently lower the pH of the local sediment environment magnification of imaging inside the flocs (Figure 6C). The size
enough to obtain polymer with increased hydrophobicity, as of aggregates formed after the degradation of poly-
inferred by the reduction in CST measurements. On the other (PCL2ChMA) significantly increased to 150 μm × 50 μm.
hand, in the case of poly(TMAEMC) and poly(TMAEMC- (Figure 6D) Recently, Zhu et al. observed similar floc
AM) the hydrophobicity of the expected hydrolysis products morphology using dual polymer system under cryo-SEM.41
under acidic conditions, poly(meth)acrylic, should not be We observed similar structures in various flocs of varying size
significantly different than that of the starting material. Hence, taken from the sediment, as shown in Figure S7.
no significant reduction in CST after the accelerated 3.6. Significance of the Current Work and Environ-
degradation study was measured for TMAEMC-based poly- mental Implication. This work addressed the challenge of
mers. efficient dewatering of MFT produced from oil sands
We analyzed our hypothesis of compaction of sediments due extraction. The composition and water chemistry of MFT
to degradation-induced hydrophobicity using cryo-FE-SEM. differs from other wastewater or industrial tailings due to the
Images of raw MFT are compared to the sediments obtained presence of residual bitumen in the mixture of kaolinite, Illite,
after the flocculation and after the degradation of 5000 ppm of and interstratified Illite-smectite and kaolinite-smectite clays.3
poly(PCL2ChMA) in Figure 6. The untreated MFT exhibited a Our group has previously attempted to treat MFT using
36297 DOI: 10.1021/acsami.7b10302
ACS Appl. Mater. Interfaces 2017, 9, 36290−36300
ACS Applied Materials & Interfaces Research Article

flocculants designed for other applications such as heavy metal polymerization of polycaprolactone choline iodide ester
removal, river sludge treatment, flocculation of microalgae and methacrylate (PCL2ChMA) macromonomers made comb-like
E. coli, quartz flocculation, etc.22,37,43−45 However, these polymeric flocculants with tunable charge density and hydro-
investigations have led us to the conclusion that it is necessary lytically degradable grafts that increased the hydrophobicity of
to design flocculants specifically tuned to balance the the polymer with under accelerated aging, further dewatering
requirements of rapid settling, recovery of water with low MFT sediments. Poly(PCL2ChMA) flocculant performed best
turbidity, and ease of dewaterability of the recovered MFT in 5 wt % MFT when added at a dosage of 5000 ppm, with an
sediment. Additionally, as MFT composition and physical ISR of 0.31 m/h, turbidity of 6.65 NTU, and CST (before
characteristics vary depending on the bitumen extraction degradation) of 22 s. The equivalent nondegradable material
process and location, it is beneficial to develop a synthesis used for comparison, poly(TMAEMC), had slightly improved
strategy that can be adapted to systematically control flocculant properties at a lower level of 3000 ppm−an ISR of 0.34 m/h,
composition and physical characteristics by tuning its hydro- turbidity of 0.405 NTU, and the lowest CST (before
phobic content and charge density. degradation) of 10.5 s−due to its higher charge density.
In particular, the PCL2ChMA polymers produced in this However, the CST of poly(PCL2ChMA) decreased by a factor
work show promise for the long-term dewatering of MFT of 2 after degradation, while that of poly(TMAEMC) did not
sediments, as the increased hydrophobicity that results from change. Copolymerization with acrylamide could control the
hydrolytic degradation of the polyester groups (previously performance of the flocculants, although higher dosages were
demonstrated for this polymer only with kaolin25) is a needed to achieve the same performance as the corresponding
successful strategy for increasing the dewaterability (and thus homopolymers.
potential land reclamation) of oil sands sediments. The novel The study demonstrates that the dewatering efficiency of the
feature of this material is that time is the trigger used to alter novel PCL2ChMA flocculants, specifically designed for treat-
the properties of the polymer, a much cheaper and easier to ment of oil sands MFT, improved upon in situ hydrolytic
implement solution than external triggers previously considered degradation. Further work is underway to relate the perform-
such as changes in temperature or pH, CO2 bubbling, and UV ance of the flocculants to the structure of the macromonomer,
exposure. Although we accelerated the degradation process in and to proceed with larger-scale testing.
this study by holding the sediment at 85 °C for 1 week,
hydrolysis of similar polyester based materials is known to
occur at ambient temperature, albeit over a longer time scale.26

*
ASSOCIATED CONTENT
S Supporting Information
Thus, hydrolytic degradation provides a simple method to The Supporting Information is available free of charge on the
improve consolidation of the tailings sediments naturally, ACS Publications website at DOI: 10.1021/acsami.7b10302.
without the need for extra energy or equipment to be employed
on the field. Mudline heights, solid content of sediments, CSTs,
In recent years, researchers have demonstrated that microbial NMR with integrations, GPC trace, FBRM, and
communities in oil sands tailings can degrade the flocculant additional SEMs (PDF)
through a methanogenic process that may produce harmful
monomers and oligomers from the polymer backbone.46,47
Though there are currently few investigations on this topic, it is
■ AUTHOR INFORMATION
Corresponding Authors
expected that further studies will emerge. Thus, it is useful to *E-mail: jsoares@ualberta.ca.
note that the expected degradation products from hydrolysis of *E-mail: robin.hutchinson@queensu.ca.
our flocculant’s PCL pendants, choline iodide and ω-hydroxy
polycaproic acid oligomers,25 are widely regarded as biode- ORCID
gradable, with similar materials being considered for biomedical Sarang P. Gumfekar: 0000-0001-7763-5360
applications.48 Both degradation products can be metabolized Author Contributions

by cellular mechanisms such as Krebb’s cycle or enzymatic S.P.G. and T.R.R. contributed equally
action.49 Furthermore, the polymer concentrations required are Notes
lower (5000 ppm relative to MFT solids) relative to other The authors declare no competing financial interest.


cationic flocculants. Considering the promising advantages of
our flocculants, we will pursue further investigations of ACKNOWLEDGMENTS
systematically modifying its structure to optimize performance,
first by changing the polyester type from polycaprolactone to The authors thank Mr. Ikenna Ezenwajiaku in the Hutchinson
polylactic acid (PLA) which hydrolytically degrades more group for the preparation of the TMAEMC-based polymers, as
rapidly, and thus would be more suitable for field conditions well as the Natural Sciences and Engineering Research Council
assuming equivalent settling performance. Second, we will of Canada (RAH group) and the Campus Alberta Innovation
Program (JBPS group) for financial support.


prepare the acrylate analogs of the macromonomers, with this
modification expected to yield flocculants of significantly higher
MW, and thus potentially faster settling rates. After the REFERENCES
relationship between optimal structure and field conditions has (1) Masliyah, J.; Zhou, Z. J.; Xu, Z.; Czarnecki, J.; Hamza, H.
been established, larger scale testing is required to proceed Understanding Water-Based Bitumen Extraction from Athabasca Oil
toward field application. Sands. Can. J. Chem. Eng. 2004, 82, 628−654.
(2) Kaminsky, H. A. W. Characterization of an Athabasca Oil Sand
Ore and Process Streams; University of Alberta: Edmonton, AB,
4. CONCLUSIONS
Canada, 2008.
In this work, we have used a new family of cationic polyester- (3) Botha, L.; Soares, J. B. P. The Influence of Tailings Composition
based polymers to flocculate oil sands MFTs. The radical on Flocculation. Can. J. Chem. Eng. 2015, 93 (9), 1514−1523.

36298 DOI: 10.1021/acsami.7b10302


ACS Appl. Mater. Interfaces 2017, 9, 36290−36300
ACS Applied Materials & Interfaces Research Article

(4) Masliyah, J. H, Czarnecki, J. A, Xu, Z. Handbook on Theory and (22) Franks, G. V. Stimulant Sensitive Flocculation and Consol-
Practice of Bitumen Recovery from Athabasca Oil Sands: Theoretical Basis, idation for Improved Solid/Liquid Separation. J. Colloid Interface Sci.
1st ed.; Kingsley Publishing Services: Alberta, Canada, 2011; Vol. 1. 2005, 292 (2), 598−603.
(5) Vajihinejad, V.; Guillermo, R.; Soares, J. B. P. Dewatering Oil (23) Chen, C. S.; Lau, Y. Y.; Mercer, S. M.; Robert, T.; Horton, J. H.;
Sands Mature Fine Tailings (MFTs) with Poly(acrylamide- co Jessop, P. G. The Effect of Switchable Water Additives on Clay
-diallyldimethylammonium chloride): Effect of Average Molecular Settling. ChemSusChem 2013, 6 (1), 132−140.
Weight and Copolymer Composition. Ind. Eng. Chem. Res. 2017, 56 (24) Eyley, S.; Vandamme, D.; Lama, S.; Van den Mooter, G.;
(5), 1256−1266. Muylaert, K.; Thielemans, W. CO2 Controlled Flocculation of
(6) Sakohara, S.; Yagi, S.; Iizawa, T. Dewatering of Inorganic Sludge Microalgae Using pH Responsive Cellulose Nanocrystals. Nanoscale
Using Dual Ionic Thermosensitive Polymers. Sep. Purif. Technol. 2011, 2015, 7 (34), 14413−14421.
80 (1), 148−154. (25) Rooney, T. R.; Gumfekar, S. P.; Soares, J. B. P.; Hutchinson, R.
(7) O’Shea, J.-P.; Qiao, G. G.; Franks, G. V. Temperature Responsive A. Cationic Hydrolytically Degradable Flocculants with Enhanced
Flocculation and Solid-Liquid Separations with Charged Random Water Recovery for Oil Sands Tailings Remediation. Macromol. Mater.
Copolymers of poly(N-isopropyl acrylamide). J. Colloid Interface Sci. Eng. 2016, 301 (10), 1248−1254.
(26) Ferrari, R.; Yu, Y.; Morbidelli, M.; Hutchinson, R. A.; Moscatelli,
2011, 360 (1), 61−70.
(8) Reis, L. G.; Oliveira, R. S.; Palhares, T. N.; Spinelli, L. S.; Lucas, D. Epsilon-Caprolactone-Based Macromonomers Suitable for Bio-
degradable Nanoparticles Synthesis Through Free Radical Polymer-
E. F.; Vedoy, D. R. L.; Asare, E.; Soares, J. B. P. Using Acrylamide/
ization. Macromolecules 2011, 44 (23), 9205−9212.
Propylene Oxide Copolymers to Dewater and Densify Mature Fine
(27) Colombo, C.; Dragoni, L.; Gatti, S.; Pesce, R. M.; Rooney, T. R.;
Tailings. Miner. Eng. 2016, 95, 29−39. Mavroudakis, E.; Ferrari, R.; Moscatelli, D. Tunable Degradation
(9) Li, Y.; Kwak, J. C. T. Rheology of Hydrophobically Modified Behavior of PEGylated Polyester-based Nanoparticles Obtained
Polyacrylamide-Co-Poly(Acrylic Acid) on Addition of Surfactant and Through Emulsion Free Radical Polymerization. Ind. Eng. Chem. Res.
Variation of Solution pH. Langmuir 2004, 20 (12), 4859−4866. 2014, 53 (22), 9128−9135.
(10) Botha, L.; Davey, S.; Nguyen, B.; Swarnakar, A. K.; Rivard, E.; (28) Franks, G. V.; Li, H.; O’Shea, J.-P.; Qiao, G. G. Temperature
Soares, J. B. P. Flocculation of Oil Sands Tailings by Hyperbranched Responsive Polymers as Multiple Function Reagents in Mineral
Functionalized Polyethylenes (HBfPE). Miner. Eng. 2017, 108, 71−82. Processing. Adv. Powder Technol. 2009, 20 (3), 273−279.
(11) Isik, M.; Fernandes, A. M.; Vijayakrishna, K.; Paulis, M.; (29) Li, H.; Long, J.; Xu, Z.; Masliyah, J. Flocculation of Kaolinite
Mecerreyes, D. Preparation of Poly(Ionic Liquid) Nanoparticles and Clay Suspensions Using a Temperature-Sensitive Polymer. AIChE J.
Their Novel Application as Flocculants for Water Purification. Polym. 2007, 53 (2), 479−488.
Chem. 2016, 7 (8), 1668−1674. (30) Mei, Y.; Ballauff, M. Effect of Counterions on The Swelling of
(12) Zheng, H.; Sun, Y.; Guo, J.; Li, F.; Fan, W.; Liao, Y.; Guan, Q. Spherical Polyelectrolyte Brushes. Eur. Phys. J. E: Soft Matter Biol. Phys.
Characterization and Evaluation of Dewatering Properties of PADB, a 2005, 16 (3), 341−349.
Highly Efficient Cationic Flocculant. Ind. Eng. Chem. Res. 2014, 53 (7), (31) Tian, B.; Ge, X.; Pan, G.; Luan, Z. Effect of Nitrate or Sulfate on
2572−2582. Flocculation Properties of Cationic Polymer Flocculants. Desalination
(13) Li, J.; Jiao, S.; Zhong, L.; Pan, J.; Ma, Q. Optimizing Coagulation 2007, 208 (1−3), 134−145.
and Flocculation Process for Kaolinite Suspension with Chitosan. (32) Alamgir, A.; Harbottle, D.; Masliyah, J.; Xu, Z. Al-PAM Assisted
Colloids Surf., A 2013, 428, 100−110. Filtration System for Abatement of Mature Fine Tailings. Chem. Eng.
(14) Sakohara, S.; Kawachi, T.; Gotoh, T.; Iizawa, T. Consolidation Sci. 2012, 80, 91−99.
of Suspended Particles by Using Dual Ionic Thermosensitive Polymers (33) Wang, C.; Han, C.; Lin, Z.; Masliyah, J.; Liu, Q.; Xu, Z. Role of
with Incorporated a Hydrophobic Component. Sep. Purif. Technol. Preconditioning Cationic Zetag Flocculant in Enhancing Mature Fine
2013, 106, 90−96. Tailings Flocculation. Energy Fuels 2016, 30 (7), 5223−5231.
(15) Zhu, Z.; Jian, O.; Paillet, S.; Desbrières, J.; Grassl, B. (34) Lu, Q.; Yan, B.; Xie, L.; Huang, J.; Liu, Y.; Zeng, H. A Two-Step
Hydrophobically Modified Associating Polyacrylamide (HAPAM) Flocculation Process on Oil Sands Tailings Treatment Using
Synthesized By Micellar Copolymerization at High Monomer Oppositely Charged Polymer Flocculants. Sci. Total Environ. 2016,
Concentration. Eur. Polym. J. 2007, 43 (3), 824−834. 565, 369−375.
(16) Deng, Y.; Xiao, H.; Pelton, R. Temperature-Sensitive (35) Lu, H.; Wang, Y.; Li, L.; Kotsuchibashi, Y.; Narain, R.; Zeng, H.
Flocculants Based on Poly(N-isopropylacrylamide-co-diallyldimethy- Temperature- and pH-Responsive Benzoboroxole-Based Polymers for
lammonium Chloride). J. Colloid Interface Sci. 1996, 179 (1), 188−193. Flocculation and Enhanced Dewatering of Fine Particle Suspensions.
(17) O’Shea, J.-P.; Qiao, G. G.; Franks, G. V. Temperature ACS Appl. Mater. Interfaces 2015, 7 (49), 27176−27187.
Responsive Flocculation and Solid-Liquid Separations with Charged (36) McClements, D. J. Theoretical Analysis of Factors Affecting the
Formation and Stability of Multilayered Colloidal Dispersions.
Random Copolymers of poly(N-isopropyl acrylamide). J. Colloid
Langmuir 2005, 21 (21), 9777−9785.
Interface Sci. 2011, 360 (1), 61−70.
(37) Feng, B.; Peng, J.; Zhu, X.; Huang, W. The Settling Behavior of
(18) Sand, A.; Yadav, M.; Mishra, D. K.; Behari, K. Modification of
Quartz Using Chitosan As Flocculant. J. Mater. Res. Technol. 2017, 6
Alginate By Grafting of N-Vinyl-2-Pyrrolidone and Studies of
(1), 71−76.
Physicochemical Properties In Terms of Swelling Capacity, Metal- (38) Das, R.; Ghorai, S.; Pal, S. Flocculation Characteristics of
Ion Uptake and Flocculation. Carbohydr. Polym. 2010, 80 (4), 1147− Polyacrylamide Grafted Hydroxypropyl Methyl Cellulose: An Efficient
1154. Biodegradable Flocculant. Chem. Eng. J. 2013, 229, 144−152.
(19) Pinaud, J.; Kowal, E.; Cunningham, M.; Jessop, P. 2- (39) Liu, Y.; Lv, C.; Ding, J.; Qian, P.; Zhang, X.; Yu, Y.; Ye, S.; Chen,
(Diethyl)aminoethyl Methacrylate as a CO2 -Switchable Comonomer Y. The Use of the Organic-Inorganic Hybrid Polymer Al(OH)3-
for the Preparation of Readily Coagulated and Redispersed Polymer Polyacrylamide to Flocculate Particles in the Cyanide Tailing
Latexes. ACS Macro Lett. 2012, 1 (9), 1103−1107. Suspensions. Miner. Eng. 2016, 89, 108−117.
(20) Zheng, H.; Sun, Y.; Zhu, C.; Guo, J.; Zhao, C.; Liao, Y.; Guan, (40) Sawalha, O.; Scholz, M. Assessment of Capillary Suction Time
Q. UV-initiated Polymerization of Hydrophobically Associating (CST) Test Methodologies. Environ. Technol. 2007, 28 (12), 1377−
Cationic Flocculants: Synthesis, Characterization, and Dewatering 1386.
Properties. Chem. Eng. J. 2013, 234, 318−326. (41) Zhu, Y.; Tan, X.; Liu, Q. Dual Polymer Flocculants for Mature
(21) Li, H.; O’Shea, J.-P.; Franks, G. V. Effect of Molecular Weight of Fine Tailings Dewatering. Can. J. Chem. Eng. 2017, 95 (1), 3−10.
Poly(N -Isopropyl Acrylamide) Temperature-Sensitive Flocculants on (42) Alagha, L.; Wang, S.; Yan, L.; Xu, Z.; Masliyah, J. Probing
Dewatering. AIChE J. 2009, 55 (8), 2070−2080. Adsorption of Polyacrylamide-Based Polymers on Anisotropic Basal

36299 DOI: 10.1021/acsami.7b10302


ACS Appl. Mater. Interfaces 2017, 9, 36290−36300
ACS Applied Materials & Interfaces Research Article

Planes of Kaolinite Using Quartz Crystal Microbalance. Langmuir


2013, 29 (12), 3989−3998.
(43) Zhang, Z.; Xia, S.; Zhang, J. Enhanced Dewatering of Waste
Sludge with Microbial Flocculant TJ-F1 as a Novel Conditioner. Water
Res. 2010, 44 (10), 3087−3092.
(44) Sharma, B. R.; Dhuldhoya, N. C.; Merchant, U. C.
Flocculantsan Ecofriendly Approach. J. Polym. Environ. 2006, 14
(2), 195−202.
(45) Bharti, S.; Mishra, S.; Sen, G. Ceric Ion Initiated Synthesis of
Polyacrylamide Grafted Oatmeal: Its Application as Flocculant for
Wastewater Treatment. Carbohydr. Polym. 2013, 93 (2), 528−536.
(46) Penner, T. J.; Foght, J. M. Mature Fine Tailings From Oil Sands
Processing Harbour Diverse Methanogenic Communities. Can. J.
Microbiol. 2010, 56 (6), 459−470.
(47) Siddique, T.; Mohamad Shahimin, M. F.; Zamir, S.; Semple, K.;
Li, C.; Foght, J. M. Long-Term Incubation Reveals Methanogenic
Biodegradation of C5 and C6 iso -Alkanes in Oil Sands Tailings.
Environ. Sci. Technol. 2015, 49 (24), 14732−14739.
(48) Agostini, A.; Gatti, S.; Cesana, A.; Moscatelli, D. Synthesis and
Degradation Study of Cationic Polycaprolactone-Based Nanoparticles
for Biomedical and Industrial Applications. Ind. Eng. Chem. Res. 2017,
56 (20), 5872−5880.
(49) Shah, A. A.; Hasan, F.; Hameed, A.; Ahmed, S. Biological
Degradation of Plastics: A Comprehensive Review. Biotechnol. Adv.
2008, 26 (3), 246−265.

36300 DOI: 10.1021/acsami.7b10302


ACS Appl. Mater. Interfaces 2017, 9, 36290−36300

You might also like