<<

metals

Article Reducing the Magnesium Content from Seawater to Improve Tailing : Description by Population Balance Models

Gonzalo R. Quezada 1 , Matías Jeldres 2,3, Norman Toro 4, Pedro Robles 5 and Ricardo I. Jeldres 3,*

1 Water Research Center for Agriculture and Mining (CRHIAM), Concepción 4030000, Chile; [email protected] 2 Faculty of Engineering and Architecture, Universidad Arturo Prat, Almirante Juan José Latorre 2901, Antofagasta 1244260, Chile; [email protected] 3 Departamento de Ingeniería Química y Procesos de Minerales, Facultad de Ingeniería, Universidad de Antofagasta, Av. Angamos 601, Antofagasta 1240000, Chile 4 Departamento de Ingeniería Metalúrgica y Minas, Universidad Católica del Norte, Antofagasta 1270709, Chile; [email protected] 5 Escuela de Ingeniería Química, Pontificia Universidad Católica de Valparaíso, Valparaíso 2340000, Chile; [email protected] * Correspondence: [email protected]; Tel.: +56-552-637-901

 Received: 18 January 2020; Accepted: 24 February 2020; Published: 2 March 2020 

Abstract: Experimental assays and mathematical models, through population balance models (PBM), were used to characterize the aggregation of mining tailings flocculated in seawater. Three systems were considered for preparation of the slurries: i) Seawater at natural pH (pH 7.4), ii) seawater at pH 11, and iii) treated seawater at pH 11. The treated seawater had a reduced magnesium content in order to avoid the formation of solid complexes, which damage the concentration operations. For this, the pH of seawater was raised with lime before being used in the process—generating solid precipitates of magnesium that were removed by vacuum filtration. The mean size of the aggregates were represented by the mean chord length obtained with the Focused beam reflectance measurement (FBRM) technique, and their descriptions, obtained by the PBM, showed an aggregation and a breakage kernel had evolved. The fractal dimension and permeability were included in the model in order to improve the representation of the irregular structure of the aggregates. Then, five parameters were optimized: Three for the aggregation kernel and two for the breakage kernel. The results show that raising the pH from 8 to 11 was severely detrimental to the flocculation performance. Nevertheless, for pH 11, the aggregates slightly exceeded 100 µm, causing undesirable behaviour during the thickening operations. Interestingly, magnesium removal provided a suitable environment to perform the tailings flocculation at alkaline pH, making aggregates with sizes that exceeded 300 µm. Only the fractal dimension changed between pH 8 and treated seawater at pH 11—as reflected in the permeability outcomes. The PBM fitted well with the experimental data, and the parameters showed that the aggregation kernel was dominant at all-polymer dosages. The descriptive capacity of the model might have been utilized as a support in practical decisions regarding the best-operating requirements in the flocculation of copper tailings and water clarification.

Keywords: copper tailings; enhanced flocculation; water recovering; magnesium removal; population balance model; seawater flocculation

Metals 2020, 10, 329; doi:10.3390/met10030329 www.mdpi.com/journal/metals Metals 2020, 10, 329 2 of 16

1. Introduction High-salinity resources, such as seawater, are being used in the mining industry in countries such as Australia, Chile, and Indonesia. At first, saltwater was proceeded by desalination plants, primarily through reverse osmosis (RO). Unfortunately, RO has created the new challenge of concentrated brine by-product, which is released into the environment; the effluent has a high temperature and a high demand for electricity, which involves fossil fuel consumption. Therefore, the direct use of seawater without ion removal is an attractive strategy that has recently been implemented in many mining industries. In this matter, engineers and researchers are constantly studying certain operational complexities. Some of these studies focus on the corrosion and incrustation of pipelines, the precipitate complexes, and the buffer effect of seawater. An example is the concentration of Cu/Mo ores, where a high level of pyrite is found, which drastically hinders the efficiency of the flotation process. The depression of pyrite in freshwater lime is used to bring the operation to an alkaline pH condition (>10.5), where the hydrophilic ferric hydroxide is formed. Nevertheless, this condition cannot function in seawater because the formation of solid Ca/Mg complexes strongly decreases the recovery of molybdenite [1,2] and chalcopyrite [3,4]. There is a debate as to the proper explanation of these findings in the sulphide flotation [5–8], but the recovery of water for the sustainability of the mining industry is more important. This is necessary due to the high cost of transporting seawater to mining sites; in Chile, the mining site can be as much as 4000 m above sea level. An essential process in the recovery of water is the thickening stage, where efficiency is linked to the flocculation of the suspended , which promotes solid–liquid separation. The flocculants are long water-soluble that can bridge multiple particles and form aggregates that are sedimentary because of the effects of gravity. The conventional flocculant used in copper mining is hydrolyzed polyacrylamide (HPAM) with a high molecular weight, which gives high settling rates, low densities, and proper rheological properties of concentrated underflows at relatively low doses and low cost [9,10]. Several efforts have been made in the study of HPAM behaviour in the solution [11,12] and high-saline conditions [13–15]. These studies provide enough proof that bridging flocculation may benefit from a saline medium because ions can act as a bridge between the anionic solid particles and anionic flocculants, commonly known as ion binding. However, the concentrated salt can reduce the size of the reagents by an excess of charge neutralization. These studies were performed on chloride salt solutions; seawater has a complex mix of ions, where Ca and Mg are present, and as we know, at high-alkaline conditions, they precipitate into a hydroxide complex. Recent studies show that kaolinite can adsorb several heavy metal ions [16,17] and also hydroxide complexes [18]; therefore, these precipitates have a strong interaction with solid particles and possibly, the flocculants. These results provide evidence that precipitates can affect the flocculation mechanism and negatively affect the thickener performance. However, this might be avoided by removing magnesium ions prior to incorporating the seawater in the operations. A good strategy to analyze flocculation performance is by monitoring the aggregate size distribution over time by the focused beam reflectance measurement (FBRM), which has the advantage of being employed directly to the slurry without sampling or the need for dilution [19–22]. In flocculated systems, a maximum floc size is achieved after a short time following flocculant addition; it then follows a decay that emerges from the fragmentation of the aggregates and polymer depletion. Using the FBRM has provided an excellent method to analyze flocculation parameters such as the maximum aggregation size or fragmentation rate [23–26]. In this context, many researchers have described the kinetics of aggregate growth and fragmentation over time by population balance models (PBM) [27–30], which have practical applications in a wide variety of systems [31–35]. The PBM is based on the work by Smoluchowski [27,36], which describes how aggregates evolve over time. Such evolution depends on the mechanisms of the aggregation and the rupture of the equations. To update the PBM to flocculation, it is necessary to improve the physical description of the aggregates by (1) updating the collision efficiency to a time-dependent flocculant depletion [30,35], and (2) to use the fractal dimension as an indicator of the irregular structure of the aggregates [27,36–40]. Recently, we Metals 2020, 10, x FOR PEER REVIEW 3 of 16

In this work, we use PBM to describe the flocculation kinetics of tailing particles using treated seawater with the magnesium removed in order to improve the water recovery and the performance of the thickening stages. Then, we analyze the implications of precipitates being present for Metalsphenomena2020, 10, 329 such as collision frequency, collision efficiency, fragmentation rate, and3 offloc 16 permeability. The outcomes of this work are of particular interest to mining industries that use seawater in concentration stages, and the implementation of this plan might allow for sustainable have proven that, for these flocculated systems, it is possible to use a constant fractal dimension when results without the need to totally remove seawater salts by reverse osmosis. 1 the shear rate is below 200 s− . 2. MethodologyIn this work, we use PBM to describe the flocculation kinetics of tailing particles using treated seawater with the magnesium removed in order to improve the water recovery and the performance of the2.1. thickening Materials stages. Then, we analyze the implications of precipitates being present for phenomena such as collision frequency, collision efficiency, fragmentation rate, and floc permeability. The outcomes The seawater (SW) was obtained from the San Jorge Bay in Antofagasta (Chile); to eliminate of this work are of particular interest to mining industries that use seawater in concentration stages, bacterial activity, the SW was filtered at 1 µm using a UV filter system. This water had a conductivity and the implementation of this plan might allow for sustainable results without the need to totally of 50.4 mS/cm, while its ionic composition was the following: 10.9 g/L Na+, 1.38 g/L Mg2+, 0.4 g/L Ca2+, remove seawater salts by reverse osmosis. 0.39 K+, 19.6 Cl−, and 0.15 g/L HCO3− [41]. 2. MethodologyKaolin was acquired from Ward’s Science, and a quantitative X-ray diffraction (XRD) analysis showed that it contained 84 wt% kaolinite (Al2Si2O5(OH)4) and 16 wt% halloysite 2.1.(Al Materials2Si2O5(OH)4·2H2O) (Figure 1). A D5000 X-ray diffractometer (Siemens S.A., Lac Condes, Chile) was usedThe and seawater the data (SW) were was processed obtained with from Total the Patte Sanrn Jorge Analysis Bay inSoftware Antofagasta (TOPAS) (Chile); (Siemens to eliminate S.A., Lac bacterialCondes, activity, Chile). theQuartz SW waswas filteredacquired at from 1 µm a using locala Chilean UV filter store, system. where This the water SiO2 hadcontent a conductivity detected by ofquantitative 50.4 mS/cm, XRD while was its ionic over composition 99 wt% (see wasFigure the 2). following: Both quartz 10.9 gand/L Na kaolin+, 1.38 had g/L a Mgdensity2+, 0.4 of g /2.6L Ca g/t.2+ ,A Microtrac+ S3500 laser diffraction particle size analyzer (Verder Scientific, Newtown, PA, USA) was 0.39 K , 19.6 Cl−, and 0.15 g/L HCO3− [41]. used.Kaolin The analysis was acquired showed from that 10% Ward’s of the Science, particles and were a smaller quantitative than d10 X-ray = 1.8 di andffraction 3.8 µm (XRD) in the kaolin and quartz samples, respectively. analysis showed that it contained 84 wt% kaolinite (Al2Si2O5(OH)4) and 16 wt% halloysite (Al Si SNFO (OH) 704, provided2H O) (Figure by SNF1). Chile A D5000 S.A., X-ray was used di ffractometer as an anionic (Siemens flocculant. S.A., The Lac molecular Condes, Chile)weight 2 2 5 4· 2 waswas used higher and than the data18 × were106. An processed initial stock with solution Total Pattern was prepared Analysis at Software a concentration (TOPAS) of (Siemens 1 g/L. Then, S.A., it was mixed with a low RPM for 24 h before use, stored in a refrigerator, and discarded after two weeks Lac Condes, Chile). Quartz was acquired from a local Chilean store, where the SiO2 content detected byin quantitative order to avoid XRD the was potential over 99 wt%ageing (see effect. Figure An2). aliquot Both quartz of this and stock kaolin flocculant had a density solution of was 2.6 gdiluted/t. A Microtracat 0.1 g/L S3500 once a laser day diforff ractionuse in testing, particle with size unused analyzer portions, (Verder then Scientific, discarded. Newtown, The flocculant PA, USA) dosages was used.were The determined analysis showedin terms thatof grams 10% of of the polymer particles per were ton of smaller dry solids than d10(g/ton).= 1.8 The and reagents 3.8 µm inused the to kaolinmodify and the quartz pH were samples, lime respectively.and sodium hydroxide and were of analytical grade (greater than 98%).

500

400

300

200 Intensity [cps]

100

0 0 20406080 Angular position [°]

FigureFigure 1. 1.X-ray X-ray di diffractionffraction (XRD) (XRD) for for kaolin kaolin powder. powder. SNF 704, provided by SNF Chile S.A., was used as an anionic flocculant. The molecular weight was higher than 18 106. An initial stock solution was prepared at a concentration of 1 g/L. Then, it × was mixed with a low RPM for 24 h before use, stored in a refrigerator, and discarded after two weeks in order to avoid the potential ageing effect. An aliquot of this stock flocculant solution was diluted at 0.1 g/L once a day for use in testing, with unused portions, then discarded. The flocculant dosages were determined in terms of grams of polymer per ton of dry solids (g/ton). The reagents used to modify the pH were lime and sodium hydroxide and were of analytical grade (greater than 98%). Metals 2020, 10, x FOR PEER REVIEW 4 of 16 Metals 2020, 10, 329 4 of 16

6000

5000

4000

3000

Intensity [cps]Intensity 2000

1000

0 0 20406080 Angular position [°] Figure 2. X-ray diffraction (XRD) for quartz powder. Figure 2. X-ray diffraction (XRD) for quartz powder. 2.2. Magnesium Removal 2.2. Magnesium Removal The removal of magnesium was carried out by increasing the pH of the seawater with 0.06 M of lime inThe order removal to form of hydroxide magnesium magnesium was carried as aout white by increa precipitate.sing the In pH parallel, of the bicarbonate seawater with was 0.06 reacted M of withlime Mg into order form to magnesium form hydroxide carbonate magnesium [41]. After as 30 a min white of intense precipitate. mixing, In the parallel, seawater bicarbonate was vacuum was filtered,reacted obtaining with Mg seawaterto form magnesium with an ionic carbonate concentration [41]. After of 10.9 30 min g/L Naof intense+, 0.01 gmixing,/L Mg2+ the, 2.35 seawater g/L Ca 2was+, vacuum+ filtered, obtaining seawater with an ionic concentration of 10.9 g/L Na+, 0.01 g/L Mg2+, 2.35 0.39 K , 19.6 Cl−, and 0.05 g/L HCO3. g/L Ca2+, 0.39 K+, 19.6 Cl−, and 0.05 g/L HCO3. 2.3. Flocculant- 2.3. Flocculant-Suspension A suspension of 270 g was prepared at 8 wt%, with known masses of the solid phases to give mixturesA suspension containing of 80 270 wt% g quartzwas prepared and 20 wt%at 8 wt%, kaolin. with The known suspension masses was of the vigorously solid phases mixed to forgive 30mixtures min using containing a 30 mm-diameter 80 wt% quartz turbine and 20 type wt% stirrer kaolin. within The suspension a 100 mm-diameter was vigorously vessel mixed with a for 1L 30 capacity.min using All a the 30 mm-diameter experiments were turbine made type by stirrer placing within the stirrer a 100 mm-diameter 20 mm above thevessel bottom with ofa 1 the L capacity. vessel. Subsequently,All the experiments the mixing were rate wasmade controlled by placing at 250 the rpm, stirrer and 20 the mm volume above of thethe solution bottom (seawater of the vessel. and polymer)Subsequently, was added the mixing in a proportion rate was fixedcontrolled by the at required 250 rpm, polymer and the dosage. volume of the solution (seawater and polymer) was added in a proportion fixed by the required polymer dosage. 2.4. Batch Settling Tests 2.4.These Batch testsSettling were Tests conducted, after 30 s of flocculant-suspension mixing, by gently pouring the slurry intoThese closed, tests 300were cm conducted,3 cylinders after (35 mm 30 internals of floc diameter),culant-suspension and then mixing, slowly invertingby gently the pouring cylinder the twoslurry times into by handclosed, (the 300 whole cm3 cylindercylinders rotation (35 mm process internal took, diameter), in all cases, and about then 4slowly s). After inverting 10 min ofthe settling,cylinder the two supernatant times by hand fluid (the was whole rescued cylinder and stirred rotation in orderprocess to homogenizetook, in all cases, the suspendedabout 4 s). After solids. 10 Then,min of a 50settling, mL aliquot the supernatant was used fluid for turbidity was rescued measurements and stirred in order a HANNA to homogenize HI98713 turbidimeterthe suspended (Hannasolids. Instruments, Then, a 50 Santiago,mL aliquot Chile), was which used performed for turbidity ten readings measurements in 20 s, delivering in a HANNA the average HI98713 at theturbidimeter end of that (Hanna period. Instruments, Santiago, Chile), which performed ten readings in 20 s, delivering the average at the end of that period. 2.5. Characterization of Aggregates 2.5.The Characterization FBRM technique of Aggregates was used to record the chord length distribution of the aggregates. The probe wasThe submergedFBRM technique vertically was in used the reactionto record vessel, the chor 10 mmd length over thedistribution stirrer and of 20 the mm aggregates. off-axis. The The FBRMprobe probe was submerged featured a laser vertically that was in the focused reaction through vessel, the 10 sapphire mm over window the stirrer and and scanned 20 mm a circular off-axis. pathThe at FBRM a tangential probe velocityfeatured of a 2laser m/s. that The backscatteredwas focused through light was the then sapphire received window when the and laser scanned beam a intersectedcircular path the pathat a tangential of the particle velocity or aggregate. of 2 m/s. AThe chord backscattered length was light determined was then from received the duration when ofthe anylaser unusual beam increaseintersected in thethe backscatteredpath of the particle light intensityor aggregate. and laserA chord velocity. length was determined from the duration of any unusual increase in the backscattered light intensity and laser velocity.

3. Modeling

Metals 2020, 10, 329 5 of 16

3. Modeling The PBM equations used in this work are derived from several works [42–44] that discretize the aggregate size into the number i of the bins. Every bin is based on the classical geometric distribution for the aggregate volume (Vi+1) = 2Vi). The PBM equation is given by:

i 2 i 1 dNi P− j (i 1) 1 P− j i dt = 2 − − Qi 1,jNi 1Nj + 2 Qi 1,j 1Ni 1Ni 1 2 − Qi,jNiNj j=1 − − − − − − − j=1 (1) imaxP1 imaxP2 Qi,jNiNj SiNi + Γi,jSjNj − j=1 − j=1 where Ni is the number concentration of the aggregates in bin i, whereas N1 is the number concentration of primary particles in bin 1. Every term on the right of the equation represents a physical process:

The first and second terms describe the aggregate formation of size i from smaller aggregates. • The third and fourth terms describe the aggregation death of size i to higher aggregates. • The fifth term represents the breakage formation of size i from the rupture of a greater aggregate. • The sixth term represents the breakage death of size i by creating smaller aggregates. • The superscript max1 is the maximum number of intervals used to represent the complete aggregate size spectrum; max2 corresponds to the largest interval from which the aggregates in the current range are produced.

3.1. Aggregation Kernel The Q variable is the aggregation kernel, an expression that contains the collision frequency (β) and capture efficiency (α): Qi,j = βi,jαi,j (2) Collision Frequency β It has been shown that fluid flow can penetrate through particle aggregates [45]. This means that the actual collision frequency is considerably lower than that predicted from rectilinear flow models. To add the permeability effects to the collision frequency, we use the parameter called “fluid collection efficiency” η. This measures the ratio between the flow passing through the aggregate and the flow approaching it. Veerapaeni and Wiesner [46] proposed a function to calculate the collision frequency, which includes the permeability and fractal dimension effects:

1 p 3 β = √η d + η d G (3) i,j 6 i i j j where, di and dj are the diameters of the aggregates sizes of i and j, respectively, G is the shear rate, and η is derived from the Brinkman’ extension to Darcy’s laws as a function of a dimensionless permeability (ξ)[47]: 9(ξi tanhξi) ηi = − (4) (3ξ + 2ξ 2 3tanhξ ) i i − i where ξi is defined as ξi = di/2 √Ki and Ki is the permeability; we use the expression from the work by Li and Logan [47]:  s  d2 i  3 3 8  K = 3 + 3 (5) i 72 1 φ − 1 φ −  − i − i The porosity φ is related to the fractal dimension (d f ) using the expression by Vainshtein et al. [48]:

!d f 3 di − φi = 1 C (6) − d0 Metals 2020, 10, 329 6 of 16

where C is a packing coefficient (generally assumed to be 0.65) and d0 the primary particle diameter. di and d0 are related by the expression proposed by Mandelbrot [49]:

1 i 1 ! d 2 f d = d − (7) i 0 C

Collision Efficiency α There are several expressions for collision efficiency, which depend on the type of aggregate and the additives. In this case, we use an expression to represent the effect of high-weight polymers; this is the polymer depletion and it rearranges on the adsorbed surface of the particles. This has been implemented in a recent work by Vajihinejad and Soares [35], which showed an exponential decay between two fitted parameters:

k t α = (αmax α )e− d + α (8) i,j − min min where αmax is the maximum collision efficiency, αmin is the minimum collision efficiency (at steady-state 1 conditions), and kd is the collision efficiency decay constant (s− ).

3.2. Breakage Kernel In regards to Equation (1), the S term is the breakage rate and Γ is the distribution breakage, which represents the breakage kernel. The S term is difficult to predict since there is no theory and is usually fitted to the size distribution data. In this case, we use a power-law function of the aggregate mass, as proposed by Pandya and Spielman [50]:

s2 Si = s1G di (9) where s1 and s2 are the fitted parameters. To simplify the distribution breakage, we use a binary distribution, where an aggregate breaks into two pieces of equal mass. That is:

 Vj  for j = i + 1  Vi Γi,j =  (10) 0 for j , i + 1

3.3. Shear Rate The shear rate required by the aggregation and breakage kernels is calculated from:

! 1 ερsus 2 G = (11) µsus where ε is the average energy dissipation rate:

3 5 NpN D ε = (12) V where Np is the impeller power number (0.6 in our case for a plane disk with gentle agitation [51]), N is the rotation speed, D and V are, respectively, the diameter of the impeller and the working volume of the vessel. The density of the suspension ρsus is calculated from:

! 1 w 1 w − ρsus = + − (13) ρs ρw Metals 2020, 10, 329 7 of 16

where w is the solid mass fraction of the solution, ρs and ρw are, respectively, the solid and water density. Finally, the viscosity of the solution µsus is measured.

3.4. Solution To resolve the PBM in Equation (1), we use a solver based on the numerical differentiation formula for stiff ODEs (ode23s) in MATLAB. The optimization was performed with the MATLAB function fminsearch, which uses the Nelder–Mead direct search to find the minimum in an unconstrained multivariable function. The objective function (OF) used to optimize the parameters is given by:

X t f  2 OF(αmax, αmin, kd, s1, s2) = dexp dmod (14) ti − where dexp is the experimental diameter of the particles, and dmod is the model diameter obtained by:

Pmax 4 i=1 Nid d = i (15) mod Pmax 3 i=1 Nidi

Two criteria validate the model’s fit and predictions: One is the coefficient of determination (R2), which measures the closeness of the model values to the experimental values. That is:

Pmax 2 i=1 dagg,exp,i dagg,mod,i R2 = 1 − (16) − Pmax 2 d dagg,exp i=1 agg,exp,i − and the other is the quality of the fit:

dagg,exp std GoF(%) = 100 − (17) dagg,exp std stands for the standard error calculated from:

1 ! 2 1 X t f  2 std = dagg,exp dagg,mod (18) n f ti − − where n is the number of data values and f is the number of parameters to be fitted. A GoF of 90% or higher means that the proposed model can predict the flocculation kinetics. Finally, conservation of the total volume of particles is verified after every integration in order to ensure that the simulations maintain the particle population.

4. Results

4.1. Input Parameters and Distribution Three cases were studied: Seawater at pH 8 (SW pH 8), seawater at pH 11 (SW pH 11), and treated seawater at pH 11 (T-SW pH 11), with Mg removed. Solving the PBM equations requires a set of input parameters (Table1). With these parameters, the shear rate can be calculated from Equations (11)–(13) and the rest are used for the PBM equation definitions. A particle number concentration is needed for each size class and for each unit of volume of suspension N0,i. For that, a number distribution is obtained from experimental results through N0,i = φv(d0,i)/V0,i, where v(d0,i) is the volume fraction of particles with diameter d0,i, obtained from Equation (7), φ is the solid volume fraction and V0,i is i 1 the volume of the primary particles following the geometric progression V0,i = 2 − V1. To obtain the v(d0,i), a volume distribution is needed. Figure3A shows an experimental volume fraction distribution provided from the FRPM probe for three case studies before the flocculant is added; in this case, all distributions are similar. To calculate the fractal dimension from the settings test, we follow the Metals 2020, 10, 329 8 of 16 procedure by Heath et al. [20]. The fractal dimension results are shown in Figure3B. The fractal dimension is higher when the flocculant dose increases. At pH 11, the fractal dimension decreases abruptly because the solid magnesium precipitates would produce open structures. Nevertheless, the treated seawater overcomes this phenomenon, giving similar structures to those of the pH 8 system.

Table 1. Input parameters and conditions.

imax 30 φ 0.054 c 0.65

Np 0.6 D 8.0 cm V 0.25 l 3 ρs 2600 kg/m 3 ρw 1000 kg/m

µsus 0.005 kg/(ms) w 0.08

Metals 2020, 10, x FOR PEER REVIEW d0 0.0005 cm 8 of 16

0.08 3.5

) SW pH 8

0,i SW pH8 f

d SW pH 11 ( 0.06 d SW pH 11 v T-SW pH 11 3.0 T-SW pH 11 0.04 2.5

2.0 0.02

Volume fraction, (A) 1.5 Fractal dimension, Fractal dimension, (B) 0.00 100 101 102 103 1.0 12 14 16 18 20 22 24 26 28 30 μ Chord length [ m], lc Flocculant dosage [g/ton] FigureFigure 3. (A 3.) (NormalizedA) Normalized initial initial volume volume distribution distribution of particle of particless of synthetic of synthetic tailings tailings in seawater in seawater at a mixing at a ratemixing of 150 raterpm. of (B 150) Fractal rpm. dimension (B) Fractal dimensionfrom the settling from the experiments. settling experiments.

4.2. Flocculation Kinetics and ModelingTable 1. Input parameters and conditions.

The results of the experimental𝑖 data are plotted 30 in Figure4 for the three cases studied in this work. We see that the flocculant doses𝜙 increase the size 0.054 of the aggregates, which is similar to other work, as shown in Reference [35]. For SW𝑐 pH 8, the aggregate 0.65 increases to a maximum value of 300 µm for the flocculant dose of 28 g/ton. Smaller𝑁 values are0.6 obtained when the flocculant dose decreases. If we increase the pH to 11 (SW pH 11),𝐷 we see results similar 8.0 to those in Figure cm 4B, where the aggregate falls to a smaller value of ~100 µm𝑉 for all the flocculant 0.25 doses in this work. l This behaviour is primarily related to the formation of hydroxide𝜌 complexes 2600 that hinder aggregatekg/m formation. Later, we see the 𝜌 1000 kg/m results of Mg removal from the seawater before being added to the slurry at pH 11 (T-SW pH 11). In 𝜇 0.005 kg/(ms) this case, we have a higher maximum aggregation than when Mg is present as a hydroxide. From this, 𝑤 0.08 we can clearly see that Mg hydroxide presence hinders the aggregation process drastically. Then, the 𝑑 0.0005 cm PBM modeling in the figures are included in the results as continuous lines. As we see from Table2, 4.2. Flocculation Kinetics and Modeling The results of the experimental data are plotted in Figure 4 for the three cases studied in this work. We see that the flocculant doses increase the size of the aggregates, which is similar to other work, as shown in Reference [35]. For SW pH 8, the aggregate increases to a maximum value of 300 µm for the flocculant dose of 28 g/ton. Smaller values are obtained when the flocculant dose decreases. If we increase the pH to 11 (SW pH 11), we see results similar to those in Figure 4B, where the aggregate falls to a smaller value of ~100 µm for all the flocculant doses in this work. This behaviour is primarily related to the formation of hydroxide complexes that hinder aggregate formation. Later, we see the results of Mg removal from the seawater before being added to the slurry at pH 11 (T-SW pH 11). In this case, we have a higher maximum aggregation than when Mg is present as a hydroxide. From this, we can clearly see that Mg hydroxide presence hinders the aggregation process drastically. Then, the PBM modeling in the figures are included in the results as continuous lines. As we see from Table 2, there is a good agreement for the cases where the mean aggregate diameter is higher, but there are limitations associated with the model for the examples that are small in size.

Metals 2020, 10, 329 9 of 16

there is a good agreement for the cases where the mean aggregate diameter is higher, but there are limitations associated with the model for the examples that are small in size. Metals 2020, 10, x FOR PEER REVIEW 9 of 16 d d 350 350 m], 13 g/ton m], 13 g/ton μ μ 300 21 g/ton 300 21 g/ton 250 28 g/ton 250 28 g/ton 200 200 150 150 100 100 50 50 (A) (B) 0 0

Mean aggregatediameter [ 0 20406080100120 Mean aggregatediameter [ 0 20406080100120

Time [s], t Time [s], t d 350 m], 13 g/ton μ 300 21 g/ton 250 28 g/ton 200 150 100 50 (C) 0

Mean aggregatediameter [ 0 20406080100120

Time [s], t FigureFigure 4. FlocculationFlocculation kineticskinetics of of synthetic synthetic tailings tailings in seawaterin seawater as a functionas a function of flocculant of flocculant doses (mixing doses (mixingrate 150 rate rpm). 150 Solid rpm). circles Solid circles correspond correspond to experimental to experimental data anddata solidand solid lines lines to the to bestthe best fit with fit with the thepopulation population balance balance model model (PBM). (PBM). (A) ( SeawaterA) Seawater (SW) (SW) pH pH 8. ( B8.) ( SWB) SW pH pH 11. 11. (C) ( TreatedC) Treated (T)-SW (T)-SW pH pH 11. 11. Table 2. Quantitative results of GoF and R2 when the PBM model is used.

TableSystem 2. Quantitative Flocculant results of Dose, GoF and g/ton R2 when theGoF PBM,% model is Rused.2 System Flocculant13 dose, g/ton 86.1GoF 0.675, % R2 SW pH 8 2113 92.986.1 0.915 0.675 28 93.8 0.939 SW pH 8 21 92.9 0.915 1328 89.993.8 0.629 0.939 SW pH 11 21 89.6 0.706 2813 91.789.9 0.794 0.629 SW pH 11 21 89.6 0.706 13 91.4 0.838 T-SW pH 11 2128 95.391.7 0.961 0.794 2813 94.991.4 0.956 0.838 T-SW pH 11 21 95.3 0.961 28 94.9 0.956

4.3. Optimized Parameters

The model fitted five parameters: The maximum and minimum collision efficiency (𝛼, 𝛼), the collision efficiency decay constant (𝑘), and two breakage rate kernel parameters (𝑠 and 𝑠). The first three parameters are related to the aggregation kernel and are shown in Figure 5. The 𝛼 show a steady increase when the flocculant dose increases. This reflects the activity of the flocculant

Metals 2020, 10, 329 10 of 16

4.3. Optimized Parameters

The model fitted five parameters: The maximum and minimum collision efficiency (αmax, αmin), the collision efficiency decay constant (kd), and two breakage rate kernel parameters (s1 and s2). The Metalsfirst three2020, 10 parameters, x FOR PEER are REVIEW related to the aggregation kernel and are shown in Figure5. The αmax10show of 16 a steady increase when the flocculant dose increases. This reflects the activity of the flocculant to to aggregate more significant quantities of solids at higher doses. Interestingly, the 𝛼 has small aggregate more significant quantities of solids at higher doses. Interestingly, the αmin has small values values compared to the 𝛼 , and all remain below 0.02, meaning that, at steady state (where 𝛼 compared to the αmax, and all remain below 0.02, meaning that, at steady state (where αmin becomes becomes significant), the breakage kernel must also be small. 𝑘 , as we found in previous results, significant), the breakage kernel must also be small. kd, as we found in previous results, must follow must follow the inverse tendency of the 𝑎 ,𝑎 because the higher the aggregates, the greater the the inverse tendency of the amax, amin because the higher the aggregates, the greater the probability probabilityof flocculant of depletion.flocculant Indepletion. this case, In the this fitted case parameters, the fitted alsoparameters show that also behaviour. show that Finally, behaviour. if we Finally,compare if thewe threecompare trials, the we three see thattrials, the we parameters see that the of parameters the SW at pH of 8the are SW similar at pH of 8 the are T-SW similar at of pH the 11, T-SWwhich at implies pH 11, that which the implies removal that of thethe Mg removal could of keep the theMg aggregation could keep performancethe aggregation similar performance to that of similarpH 8.Finally, to that of the pH two 8. parametersFinally, the fortwo the parameters breakage for kernel the arebreakage plotted kernel in Figure are plotted6. As discussed in Figure above, 6. As discussed above, if the s1 is small, the 𝑎 is also small. This trend must be followed in order to if the s1 is small, the amin is also small. This trend must be followed in order to satisfy steady-state satisfy steady-state conditions in the aggregation process. The s2 parameters show values between 1 conditions in the aggregation process. The s2 parameters show values between 1 and 2, but as the s1 andparameter 2, but as is small,the s1 parameter its contribution is small, is neglected its contribution from the is global neglected behavior. from Fromthe global this, webehavior. see that From high this,dosage we incrementssee that high contribute dosage increments to the aggregate contribute kernel to ratherthe aggregate than the kernel breakage rather kernel. than the breakage kernel. 0.10 0.05 SW pH 8 SW pH 8 0.08 SW pH 11 0.04 SW pH 11 T-SW pH 11 T-SW pH 11 0.06 0.03 min max α α 0.04 0.02

0.02 (A) 0.01 (B)

0.00 0.00 12 14 16 18 20 22 24 26 28 30 12 14 16 18 20 22 24 26 28 30

Flocculant dosage [g/ton] Flocculant dosage [g/ton] 0.4 SW pH 8 SW pH 11 0.3 T-SW pH 11 ] -1

[s 0.2 d k

0.1 (C)

0.0 12 14 16 18 20 22 24 26 28 30

Flocculant dosage [g/ton] FigureFigure 5. 5. OptimumOptimum aggregation aggregation parameters parameters vs. vs. shear shear rate rate for for constant constant and and variab variablele fractal fractal dimension: dimension:

((AA)) Maximum Maximum and and ( (BB)) minimum minimum collision collision efficiency efficiency and and ( (CC)) collision collision efficiency efficiency decay constant k𝑘d..

Metals 2020, 10, 329 11 of 16 Metals 2020, 10, x FOR PEER REVIEW 11 of 16

0.10 3.0 SW pH 8 (B) 0.08 SW pH 11 2.5 T-SW pH 11 2.0 0.06 1 2

s s 1.5 0.04 1.0 SW pH 8 0.02 (A) 0.5 SW pH 11 T-SW pH 11 0.00 0.0 12 14 16 18 20 22 24 26 28 30 12 14 16 18 20 22 24 26 28 30

Flocculant dosage [g/ton] Flocculant dosage [g/ton]

Figure 6.6. OptimumOptimum breakagebreakage parameters parameters vs. vs. shear shear rate rate for fo constantr constant and and variable variable fractal fractal dimension: dimension: (A) (A) S1 and (B) collision efficiency decay constant 𝑘 . S1 and (B) collision efficiency decay constant kd.

4.4. Aggregation, Breakage, and Permeability Modeling Once wewe havehave thethe optimizedoptimized parameters,parameters, wewe cancan computecompute thethe physicalphysical parameters:parameters: CollisionCollision eefficiency,fficiency, collisioncollision frequency, frequency, breakage breakage rate, rate, and an permeabilityd permeability with with the the mean mean aggregate aggregate size size from from the resultsthe results in Figure in Figure4. Collision 4. Collision frequencies frequencies are shown are inshown Figure in7 .Figure We see 7. that We high see dosesthat high of flocculant doses of increaseflocculant the increase collision the frequency. collision Thisfrequency. can be accountedThis can be for accounted by the polymer for by concentration,the polymer concentration, which has an extendedwhich has configuration an extended thatconfiguration increases the that likelihood increases of the collisions. likelihood For of SW collisions. pH 11, the For small SW aggregatespH 11, the generatedsmall aggregates by the appearancegenerated by of the magnesium appearance complexes of magnesium lead to acomplexes low value lead in the to collisiona low value frequency. in the Thiscollision does frequency. not present This significant does not changes present with significant varied flocculant changes doses. with Calculationsvaried flocculant for collision doses. eCalculationsfficiency are presentedfor collision in Figureefficiency8 and are follow presented the same in trendFigure as the8 and collision follow frequency, the same where trend SW as pHthe 11collision has no frequency, significant values.where SW In this pH case,11 has only no below significant 20 s is values. the parameter In this case, relevant only to below the kinetics, 20 s is andthe dropsparameter to zero relevant values to after the 20kinetics, s. Breakage and drops rates to are ze shownro values in Figure after 209, whichs. Breakage shows rates that, are at SWshown pH 8,in weFigure see that9, which low dosesshows generate that, at SW higher pH values 8, we see because that low less doses flocculant generate cannot higher maintain values the because aggregates. less Weflocculant see similar cannot behaviour maintain in the SW aggregates. pH 11 and We T-SW see pH similar 11; this behaviour is accounted in SW for pH because 11 and ofT-SW the pH fractal 11; dimensions.this is accounted For SWfor because at pH 11, of the the fractal fractal dimension dimensions. is low,For SW and at the pH diameter 11, the fractal increase dimension coincided is withlow, anand increased the diameter breakage increase rate. coincided For T-SW with at pH an 11, increased the fractal breakage dimension rate. is high,For T-SW decreasing at pH 11, the the aggregate fractal diameterdimension and is decreasinghigh, decreasing the breakage the aggregate rate, even diameter if the aggregates and decreasing are big. the Finally, breakage permeability rate, even results if the wereaggregates plotted are in big. Figure Finally, 10; we permea can seebility that results very porous were plotted aggregates in Figure are found 10; we for can SW see pH that 8 and very a doseporous of 21aggregates g/ton. This are can found be accounted for SW pH for 8 by and the a fractaldose of dimension 21 g/ton. andThis the can large be accounted mean size offor the by aggregates.the fractal Fordimension SW at pH and 11, the we large see thatmean permeability size of the isaggregates. lower even For if theSW fractal at pH dimension11, we see isthat small, permeability because the is sizelower of even the aggregate if the fractal is small dimension in these is cases. small, Interestingly, because thethe size permeability of the aggregate value is for small T-SW in pH these 11 iscases. low, asInterestingly, compared to the other permeability cases, even value if the for aggregate T-SW pH sizes 11 areis low, similar as compared to those in to SW other pH 8.cases, This even is because if the theaggregate fractal dimensionsizes are similar is high to and those creates in SW dense pH aggregates8. This is because that have the little fractal permeability. dimension is high and creates dense aggregates that have little permeability. 5 5

10 2.5 10 2.5 ii ii ii β (A) 13 g/ton β (B) 13 g/ton ], ],

-1 2.0 21 g/ton -1 2.0 21 g/ton s s 3 28 g/ton 3 28 g/ton 1.5 1.5

1.0 1.0

0.5 0.5

0.0 0.0

Collision frequency [cm 0 20406080100120 Collision frequency [cm 0 20406080100120 Time [s], t Time [s], t

Metals 2020, 10, x FOR PEER REVIEW 11 of 16

0.10 3.0 SW pH 8 (B) 0.08 SW pH 11 2.5 T-SW pH 11 2.0 0.06 1 2

s s 1.5 0.04 1.0 SW pH 8 0.02 (A) 0.5 SW pH 11 T-SW pH 11 0.00 0.0 12 14 16 18 20 22 24 26 28 30 12 14 16 18 20 22 24 26 28 30

Flocculant dosage [g/ton] Flocculant dosage [g/ton] Figure 6. Optimum breakage parameters vs. shear rate for constant and variable fractal dimension:

(A) S1 and (B) collision efficiency decay constant 𝑘.

4.4. Aggregation, Breakage, and Permeability Modeling Once we have the optimized parameters, we can compute the physical parameters: Collision efficiency, collision frequency, breakage rate, and permeability with the mean aggregate size from the results in Figure 4. Collision frequencies are shown in Figure 7. We see that high doses of flocculant increase the collision frequency. This can be accounted for by the polymer concentration, which has an extended configuration that increases the likelihood of collisions. For SW pH 11, the small aggregates generated by the appearance of magnesium complexes lead to a low value in the collision frequency. This does not present significant changes with varied flocculant doses. Calculations for collision efficiency are presented in Figure 8 and follow the same trend as the collision frequency, where SW pH 11 has no significant values. In this case, only below 20 s is the parameter relevant to the kinetics, and drops to zero values after 20 s. Breakage rates are shown in Figure 9, which shows that, at SW pH 8, we see that low doses generate higher values because less flocculant cannot maintain the aggregates. We see similar behaviour in SW pH 11 and T-SW pH 11; this is accounted for because of the fractal dimensions. For SW at pH 11, the fractal dimension is low, and the diameter increase coincided with an increased breakage rate. For T-SW at pH 11, the fractal dimension is high, decreasing the aggregate diameter and decreasing the breakage rate, even if the aggregates are big. Finally, permeability results were plotted in Figure 10; we can see that very porous aggregates are found for SW pH 8 and a dose of 21 g/ton. This can be accounted for by the fractal dimension and the large mean size of the aggregates. For SW at pH 11, we see that permeability is lower even if the fractal dimension is small, because the size of the aggregate is small in these cases. Interestingly, the permeability value for T-SW pH 11 is low, as compared to other cases, even if the Metalsaggregate2020, 10 sizes, 329 are similar to those in SW pH 8. This is because the fractal dimension is high12 ofand 16 creates dense aggregates that have little permeability. 5 5

10 2.5 10 2.5 ii ii ii β (A) 13 g/ton β (B) 13 g/ton ], ],

-1 2.0 21 g/ton -1 2.0 21 g/ton s s 3 28 g/ton 3 28 g/ton 1.5 1.5

1.0 1.0

0.5 0.5

0.0 0.0 Metals 2020, 10, x FORCollision frequency [cm PEER0 REVIEW 20406080100120 Collision frequency [cm 0 20406080100120 12 of 16 Metals 2020, 10, x FOR PEER REVIEW 12 of 16 Time [s], t Time [s], t 5

5 10 2.5 ii ii 10 2.5β ii ii (C) β 13 g/ton ], (C) 13 g/ton -1 ], 2.0 21 g/ton s -1 3 2.0 21 g/ton s

3 28 g/ton 28 g/ton 1.5 1.5 1.0 1.0 0.5 0.5 0.0 0.0 Collision frequency [cm 0 20406080100120 Collision frequency [cm 0 20406080100120 Time [s], t Time [s], t Figure 7. Collision frequency for the experimental data for (A) SW pH 8, (B) SW pH 11, and (C) T-SW FigureFigure 7.7. Collision frequency forfor thethe experimentalexperimental datadata forfor ((AA)) SWSW pHpH 8,8, ((BB)) SWSW pHpH 11,11, andand ((C)) T-SWT-SW pH 11. pHpH 11.11.

0.10 0.10 0.10 (A) 0.10 (B) (A) 13 g/ton (B) 13 g/ton α 13 g/ton α 13 g/ton α 0.08 21 g/ton α 0.08 21 g/ton 0.08 21 g/ton 0.08 21 g/ton 28 g/ton 28 g/ton 28 g/ton 28 g/ton 0.06 0.06 0.06 0.06 0.04 0.04 0.04 0.04 0.02 0.02

0.02Collision efficiency, 0.02Collision efficiency, Collision efficiency, Collision efficiency, 0.00 0.00 0.00 0 20406080100120 0.00 0 20406080100120 0 20406080100120 0 20406080100120 Time [s], t Time [s], t Time [s], t Time [s], t 0.10 0.10 (C) (C) 13 g/ton α 13 g/ton α 0.08 21 g/ton 0.08 21 g/ton 28 g/ton 28 g/ton 0.06 0.06 0.04 0.04 0.02

0.02Collision efficiency, Collision efficiency, 0.00 0.00 0 20406080100120 0 20406080100120 Time [s], t Time [s], t Figure 8. Collision efficiency for the experimental data for (A) SW pH 8, (B) SW pH 11, and (C) T-SW FigureFigure 8.8. CollisionCollision eefficiencyfficiency forfor thethe experimentalexperimental datadata forfor ((AA)) SWSW pHpH 8,8, ((BB)) SWSW pHpH 11,11, andand ((C)) T-SWT-SW pH 11. pHpH 11.11.

0.25 0.25 0.25 (A) 0.25 (B) i (A) 13 g/ton i (B) 13 g/ton S S i 13 g/ton i 13 g/ton S 0.20 21 g/ton S 0.20 21 g/ton ], ],

-1 0.20 21 g/ton -1 0.20 21 g/ton ], 28 g/ton ], 28 g/ton -1 28 g/ton -1 28 g/ton 0.15 0.15 0.15 0.15 0.10 0.10 0.10 0.10 0.05 0.05

0.05 [s rate Breakage 0.05 [s rate Breakage Breakage rate [s rate Breakage [s rate Breakage 0.00 0.00 0.00 0 20406080100120 0.00 0 20 40 60 80 100 120 0 20406080100120 0 20 40 60 80 100 120 Time [s], t Time [s], t Time [s], t Time [s], t

Metals 2020, 10, x FOR PEER REVIEW 12 of 16 5

10 2.5 ii ii β (C) 13 g/ton ],

-1 2.0 21 g/ton s 3 28 g/ton 1.5

1.0

0.5

0.0

Collision frequency [cm 0 20406080100120 Time [s], t Figure 7. Collision frequency for the experimental data for (A) SW pH 8, (B) SW pH 11, and (C) T-SW pH 11.

0.10 0.10 (A) 13 g/ton (B) 13 g/ton α α 0.08 21 g/ton 0.08 21 g/ton 28 g/ton 28 g/ton 0.06 0.06

0.04 0.04

0.02 0.02 Collision efficiency, Collision efficiency,

0.00 0.00 0 20406080100120 0 20406080100120 Time [s], t Time [s], t 0.10 (C) 13 g/ton α 0.08 21 g/ton 28 g/ton 0.06

0.04

0.02 Collision efficiency,

0.00 0 20406080100120 Time [s], t

Metals 2020Figure, 10, 329 8. Collision efficiency for the experimental data for (A) SW pH 8, (B) SW pH 11, and (C) T-SW13 of 16 pH 11.

0.25 0.25 (A) (B) i 13 g/ton i 13 g/ton S 0.20 21 g/ton S 0.20 21 g/ton ], ],

-1 28 g/ton -1 28 g/ton 0.15 0.15

0.10 0.10

0.05 0.05 Breakage rate [s rate Breakage [s rate Breakage Metals 2020, 10, x FOR PEER REVIEW 13 of 16 0.00 0.00 0 20406080100120 0 20 40 60 80 100 120 Metals 2020, 10, x FOR PEER REVIEW 13 of 16 Time [s], t 0.25 Time [s], t (C) i 13 g/ton S 0.250.20 21 g/ton ], (C) -1 i 1328 g/ton S 0.20 21 g/ton ], 0.15 -1 28 g/ton 0.150.10

0.100.05 Breakage rate [s rate Breakage

0.050.00 Breakage rate [s rate Breakage 0 20 40 60 80 100 120 0.00 Time [s], t 0 20 40 60 80 100 120 Figure 9. Breakage rate for the experimental dataTime for [s], ( At ) SW pH 8, (B) SW pH 11, and (C) T-SW pH FigureFigure11. 9. Breakage9. Breakage rate rate for for the the experimental experimental data data for (forA) SW(A) pHSW 8,pH (B )8, SW (B) pH SW 11, pH and 11, ( Cand) T-SW (C) T-SW pH 11. pH 11. 2.0 2.0 3 (A) 13 g/ton 3 (B) 13 g/ton 10 10 2.0 21 g/ton 2.0 21 g/ton 3 3 K K 1.5 (A) 13 g/ton 1.5 (B) 13 g/ton ], ], 28 g/ton ], 28 g/ton 10 10 2 21 g/ton 2 21 g/ton K K 1.5 1.5

], ], 28 g/ton ], 28 g/ton 2 1.0 2 1.0

1.0 1.0 0.5 0.5

Permeability [cm Permeability 0.5 [cm Permeability 0.5 0.0 0.0

Permeability [cm Permeability 0 20406080100120 [cm Permeability 0 20406080100120 0.0 0.0 0 20406080100120Time [s], t 0 20406080100120Time [s], t Time [s], t 2.0 Time [s], t

3 (C) 13 g/ton 10 2.0 21 g/ton 3 K 1.5 (C) 13 g/ton ], 28 g/ton 10 2 21 g/ton K 1.5

], 28 g/ton 2 1.0

1.0 0.5

Permeability [cm Permeability 0.5 0.0

Permeability [cm Permeability 0 20406080100120 0.0 0 20406080100120Time [s], t Time [s], t FigureFigure 10. 10.Permeability Permeability for for the the experimental experimental data data for (Afor) SW(A) pHSW 8, pH (B )8, SW (B) pH SW 11, pH and 11, ( Cand) T-SW (C) T-SW pH 11. pH Figure11. 10. Permeability for the experimental data for (A) SW pH 8, (B) SW pH 11, and (C) T-SW pH 5. Conclusions 11. 5. ExperimentsConclusions and modeling, through a population balance model (PBM), are used to characterize the5. particle ConclusionsExperiments aggregation and modeling, of flocculated through tailings a population on untreated balance and model treated (PBM), seawater are used (SW) to at characterize different flocculantthe particleExperiments doses. aggregation It was and foundmodeling, of flocculated that through the Mg tailings hydroxidea population on untreated complex balance presenceand model treated (PBM), could seawater hinderare used (SW) the to aggregatecharacterize at different kineticstheflocculant particle drastically doses. aggregation overIt was time; foundof flocculated this that generates the tailingsMg smallhydr onoxide aggregates untreated complex betweenand presence treated 50–100 couldseawater microns hinder (SW) atthe pHat aggregate different 11. If flocculantkinetics drastically doses. It wasover found time; thisthat generatesthe Mg hydr smaloxidel aggregates complex betweenpresence 50–100could hindermicrons the at aggregate pH 11. If kineticsthe Mg isdrastically removed beforeover time; the flocculation this generates takes smal place,l aggregates we can obtain between larger 50–100 aggregates, microns from at pH 100–200 11. If themicrons, Mg is atremoved the same before pH. the The flocculation results of takesT-SW pl pHace, 11 we are can in obtain good larger agreement aggregates, with thefrom SW 100–200 pH 8 microns,achievements, at the which same maypH. Theimply results that pHof T-SWis no pHlonger 11 area restrictive in good agreementparameter forwith the the aggregation SW pH 8 achievements,process. The modeling which may of the imply aggregation that pH was is nodone longer with aPBM, restrictive where parameterfitted parameters for the can aggregation represent process.the flocculation The modeling aggregation of the kinetics.aggregation In this was case, done th withe effect PBM, of flocculantwhere fitted dose parameters can neglect can the represent impact theof the flocculation breakage aggregation kernel, and kinetics.the aggregation In this case, kernel the effectdominates of flocculant the flocculati dose canon. neglectAdditionally, the impact it is offound the that,breakage without kernel, Mg ions,and thethe aggregationpH has little kerneleffect on dominates both the flocculationthe flocculati kineticson. Additionally, and PBM fitted it is foundparameters. that, without Only the Mg fractal ions, dimensionthe pH has showslittle effect us the on main both differencethe flocculation between kinetics SW pHand 8 PBM and T-SWfitted parameters.pH 11, where Only it is the reflected fractal indimension its permeability. shows us As the we main can see,difference experiments between and SW modeling pH 8 and show T-SW us pH 11, where it is reflected in its permeability. As we can see, experiments and modeling show us

Metals 2020, 10, 329 14 of 16 the Mg is removed before the flocculation takes place, we can obtain larger aggregates, from 100–200 microns, at the same pH. The results of T-SW pH 11 are in good agreement with the SW pH 8 achievements, which may imply that pH is no longer a restrictive parameter for the aggregation process. The modeling of the aggregation was done with PBM, where fitted parameters can represent the flocculation aggregation kinetics. In this case, the effect of flocculant dose can neglect the impact of the breakage kernel, and the aggregation kernel dominates the flocculation. Additionally, it is found that, without Mg ions, the pH has little effect on both the flocculation kinetics and PBM fitted parameters. Only the fractal dimension shows us the main difference between SW pH 8 and T-SW pH 11, where it is reflected in its permeability. As we can see, experiments and modeling show us that the Mg removal is the main component that hinders particle aggregation, and without its presence, the process can operate similarly at neutral pH and high pH.

Author Contributions: All of the authors contributed to the analysis of the results and writing the paper. All authors have read and agreed to the published version of the manuscript Funding: This research was funded by ANID Fondecyt 11171036, ANID ACM 170005, and Centro CRHIAM Project ANID/Fondap/15130015. Acknowledgments: Gonzalo Quezada and Ricardo I. Jeldres thank the Centro CRHIAM Project ANID/Fondap/15130015. Ricardo I. Jeldres thanks ANID Fondecyt 11171036 and ANID ACM 170005. Pedro Robles thanks the Pontificia Universidad Católica de Valparaíso for the support provided. Conflicts of Interest: The authors declare no conflict of interest.

References

1. Qiu, Z.; Liu, G.; Liu, Q.; Zhong, H. Understanding the roles of high salinity in inhibiting the molybdenite flotation. Surfaces A Physicochem. Eng. Asp. 2016.[CrossRef] 2. Li, W.; Li, Y.; Wei, Z.; Xiao, Q.; Song, S. Fundamental studies of SHMP in reducing negative effects of divalent ions on molybdenite flotation. Minerals 2018, 8, 404. [CrossRef] 3. Li, Y.; Li, W.; Xiao, Q.; He, N.; Ren, Z.; Lartey, C.; Gerson, A. The influence of common monovalent and divalent chlorides on chalcopyrite flotation. Minerals 2017, 7, 111. [CrossRef] 4. Uribe, L.; Gutierrez, L.; Laskowski, J.S.; Castro, S. Role of calcium and magnesium cations in the interactions between kaolinite and chalcopyrite in seawater. Physicochem. Probl. Miner. Process. 2017, 53, 737–749. [CrossRef] 5. Ramos, O.; Castro, S.; Laskowski, J.S. Copper–molybdenum ores flotation in sea water: Floatability and frothability. Miner. Eng. 2013, 53, 108–112. [CrossRef] 6. Hirajima, T.; Suyantara, G.P.W.; Ichikawa, O.; Elmahdy, A.M.; Miki, H.; Sasaki, K. Effect of Mg2+ and Ca2+ as divalent seawater cations on the floatability of molybdenite and chalcopyrite. Miner. Eng. 2016, 96–97, 83–93. [CrossRef] 7. Jeldres, R.I.; Arancibia-Bravo, M.P.; Reyes, A.; Aguirre, C.E.; Cortes, L.; Cisternas, L.A. The impact of seawater with calcium and magnesium removal for the flotation of copper-molybdenum sulphide ores. Miner. Eng. 2017, 109, 10–13. [CrossRef] 8. Suyantara, G.P.W.; Hirajima, T.; Miki, H.; Sasaki, K. Floatability of molybdenite and chalcopyrite in artificial seawater. Miner. Eng. 2018, 115, 117–130. [CrossRef] 9. Lee, L.T.; Rahbari, R.; Lecourtier, J.; Chauveteau, G. Adsorption of polyacrylamides on the different faces of kaolinites. J. Interface Sci. 1991, 147, 351–357. [CrossRef] 10. Nasser, M.S.; James, A.E. Effect of polyacrylamide polymers on floc size and rheological behaviour of kaolinite suspensions. Colloids Surfaces A Physicochem. Eng. Asp. 2007, 301, 311–322. [CrossRef] 11. Zhang, L.; Zheng, L.; Pu, J.; Pu, C.; Cui, S. Influence of hydrolyzed polyacrylamide (hpam) molecular weight on the cross-linking reaction of the HPAM/Cr3+ System and transportation of the HPAM/Cr3+ system in microfractures. Energy Fuels 2016.[CrossRef] 12. Abdel-Azeim, S.; Kanj, M.Y. Dynamics, Aggregation, and Interfacial Properties of the Partially Hydrolyzed Polyacrylamide Polymer for Enhanced Oil Recovery Applications: Insights from Molecular Dynamics Simulations. Energy Fuels 2018.[CrossRef] Metals 2020, 10, 329 15 of 16

13. Quezada, G.R.; Jeldres, R.I.; Fawell, P.D.; Toledo, P.G. Use of molecular dynamics to study the conformation of an anionic in saline medium and its adsorption on a quartz surface. Miner. Eng. 2018, 129, 102–105. [CrossRef] 14. Quezada, G.R.; Saavedra, J.H.; Rozas, R.E.; Toledo, P.G. Molecular dynamics simulations of the conformation and diffusion of partially hydrolyzed polyacrylamide in highly saline solutions. Chem. Eng. Sci. 2019. [CrossRef] 15. Jeldres, M.; Piceros, E.C.; Toro, N.; Torres, D.; Robles, P.; Leiva, W.H.; Jeldres, R.I. Copper tailing flocculation in seawater: Relating the yield stress with fractal aggregates at varied mixing conditions. Metals 2019, 9, 1295. [CrossRef] 16. Jiang, M.Q.; Jin, X.Y.; Lu, X.Q.; Chen, Z.L. Adsorption of Pb(II), Cd(II), Ni(II) and Cu(II) onto natural kaolinite clay. Desalination 2010.[CrossRef] 17. Bhattacharyya, K.G.; Gupta, S.S. Adsorption of a few heavy metals on natural and modified kaolinite and montmorillonite: A review. Adv. Colloid Interface Sci. 2008, 140, 114–131. [CrossRef] 18. Diamond, S.; Kinter, E.B. Adsorption of calcium hydroxide by montmorillonite and kaolinite. J. Colloid Interface Sci. 1966.[CrossRef] 19. Biggs, C.A.; Lant, P.A. Activated sludge flocculation: On-line determination of floc size and the effect of shear. Water Res. 2000.[CrossRef] 20. Heath, A.R.; Bahri, P.A.; Fawell, P.D.; Farrow, J.B. Polymer flocculation of calcite: Relating the aggregate size to the settling rate. AIChE J. 2006, 52, 1987–1994. [CrossRef] 21. Gregory, J. Monitoring particle aggregation processes. Adv. Colloid Interface Sci. 2009, 147–148, 109–123. [CrossRef][PubMed] 22. Meng, Z.; Hashmi, S.M.; Elimelech, M. Aggregation rate and fractal dimension of fullerene via simultaneous multiangle static and measurement. J. Colloid Interface Sci. 2013, 392, 27–33. [CrossRef][PubMed] 23. Heath, A.R.; Bahri, P.A.; Fawell, P.D.; Farrow, J.B. Polymer flocculation of calcite: Experimental results from turbulent pipe flow. AIChE J. 2006, 52, 1284–1293. [CrossRef] 24. Owen, A.T.; Fawell, P.D.; Swift, J.D.; Labbett, D.M.; Benn, F.A.; Farrow, J.B. Using turbulent pipe flow to study the factors affecting polymer-bridging flocculation of mineral systems. Int. J. Miner. Process. 2008, 87, 90–99. [CrossRef] 25. Bubakova, P.; Pivokonsky, M.; Filip, P. Effect of shear rate on aggregate size and structure in the process of aggregation and at steady state. Powder Technol. 2013, 235, 540–549. [CrossRef] 26. Benn, F.A.; Fawell, P.D.; Halewood, J.; Austin, P.J.; Costine, A.D.; Jones, W.G.; Francis, N.S.; Druett, D.C.; Lester, D. and consolidation of different density aggregates formed by polymer-bridging flocculation. Chem. Eng. Sci. 2018, 184, 111–125. [CrossRef] 27. Thomas, D.N.; Judd, S.J.; Fawcett, N. Flocculation modelling: a review. Water Res. 1999, 33, 1579–1592. [CrossRef] 28. Heath, A.R.; Bahri, P.A.; Fawell, P.D.; Farrow, J.B. Polymer flocculation of calcite: Population balance model. AIChE J. 2006, 52, 1641–1653. [CrossRef] 29. Jeldres, R.I.; Fawell, P.D.; Florio, B.J. Population balance modelling to describe the particle aggregation process: A review. Powder Technol. 2018, 326, 190–207. [CrossRef] 30. Quezada, G.R.; Ramos, J.; Jeldres, R.I.; Robles, P.; Toledo, P.G. Analysis of the flocculation process of fine tailings particles in saltwater through a population balance model. Sep. Purif. Technol. 2019, 237, 116319. [CrossRef] 31. Runkana, V.; Somasundaran, P.; Kapur, P.C. A population balance model for flocculation of colloidal suspensions by polymer bridging. Chem. Eng. Sci. 2006, 61, 182–191. [CrossRef] 32. Costa, C.B.B.; Maciel, M.R.W.; Filho, R.M. Considerations on the crystallization modeling: Population balance solution. Comput. Chem. Eng. 2007, 31, 206–218. [CrossRef] 33. Datta, A.; Rajamani, R.K. A direct approach of modeling batch grinding in ball mills using population balance principles and impact energy distribution. Int. J. Miner. Process. 2002, 64, 181–200. [CrossRef] 34. Kiparissides, C.; Alexopoulos, A.; Roussos, A.; Dompazis, G.; Kotoulas, C. Population balance modeling of particulate polymerization processes. Ind. Eng. Chem. Res. 2004, 43, 7290–7302. [CrossRef] 35. Vajihinejad, V.; Soares, J.B.P. Monitoring polymer flocculation in oil sands tailings: A population balance model approach. Chem. Eng. J. 2018, 346, 447–457. [CrossRef] Metals 2020, 10, 329 16 of 16

36. Jeldres, R.I.; Concha, F.; Toledo, P.G. Population balance modelling of particle flocculation with attention to aggregate restructuring and permeability. Adv. Colloid Interface Sci. 2015, 224, 62–71. [CrossRef] 37. Ahmad, A.L.; Chong, M.F.; Bhatia, S. Population Balance Model (PBM) for flocculation process: Simulation and experimental studies of palm oil mill effluent (POME) pretreatment. Chem. Eng. J. 2008, 140, 86–100. [CrossRef] 38. Flesch, J.C.; Spicer, P.T.; Pratsinis, S.E. Laminar and turbulent shear-induced flocculation of fractal aggregates. AIChE J. 1999, 45, 1114–1124. [CrossRef] 39. Selomulya, C.; Bushell, G.; Amal, R.; Waite, T.D. Understanding the role of restructuring in flocculation: The application of a population balance model. Chem. Eng. Sci. 2003, 58, 327–338. [CrossRef] 40. Antunes, E.; Garcia, F.A.P.; Ferreira, P.; Blanco, A.; Negro, C.; Rasteiro, M.G. Modelling PCC flocculation by bridging mechanism using population balances: Effect of polymer characteristics on flocculation. Chem. Eng. Sci. 2010, 65, 3798–3807. [CrossRef] 41. Jeldres, M.; Piceros, E.; Robles, P.A.; Toro, N.; Jeldres, R.I. Viscoelasticity of quartz and kaolin slurries in seawater: Importance of magnesium precipitates. Metals 2019, 9, 1120. [CrossRef] 42. Spicer, P.T.; Pratsinis, S.E. Shear-induced flocculation: The evolution of floc structure and the shape of the size distribution at steady state. Water Res. 1996, 30, 1049–1056. [CrossRef] 43. Hounslow, M.J.; Ryall, R.L.; Marshall, V.R. A discretized population balance for nucleation, growth, and aggregation. AIChE J. 1988, 34, 1821–1832. [CrossRef] 44. Kusters, K.A.; Pratsinis, S.E.; Smith, D.M.; Thoma, S.G. Ultrasonic fragmentation of agglomerate powders. Chem. Eng. Sci. 1993, 48, 4119–4127. [CrossRef] 45. Thill, A.; Moustier, S.; Aziz, J.; Wiesner, M.R.; Bottero, J.Y. Flocs restructuring during aggregation: Experimental evidence and numerical simulation. J. Colloid Interface Sci. 2001, 243, 171–182. [CrossRef] 46. Veerapaneni, S.; Wiesner, M.R. Hydrodynamics of fractal aggregates with radially varying permeability. J. Colloid Interface Sci. 1996, 177, 45–57. [CrossRef] 47. Li, X.-Y.; Logan, B.E. Permeability of fractal aggregates. Water Res. 2001, 35, 3373–3380. [CrossRef] 48. Vainshtein, P.; Shapiro, M.; Gutfinger, C. Mobility of permeable aggregates: effects of shape and porosity. J. Aerosol Sci. 2004, 35, 383–404. [CrossRef] 49. Mandelbrot, B.B. Self-affine fractals and fractal dimension. Phys. Scr. 1985, 32, 257–260. [CrossRef] 50. Pandya, J.D.; Spielman, L.A. Floc breakage in agitated suspensions: Effect of agitation rate. Chem. Eng. Sci. 1983, 38, 1983–1992. [CrossRef] 51. Pretorius, C.; Wicklein, E.; Rauch-Williams, T.; Samstag, R.; Sigmon, C. How oversized mixers became an industry standard. In Proceedings of the 88th Annual Water Environment Federation Technical Exhibition and Conference, WEFTEC 2015, Chicago, IL, USA, 26–30 September 2015.

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).