You are on page 1of 19

minerals

Article
Parametric Optimization in Rougher Flotation
Performance of a Sulfidized Mixed Copper Ore
Asghar Azizi 1 , Mojtaba Masdarian 1 , Ahmad Hassanzadeh 2 , Zahra Bahri 3 ,
Tomasz Niedoba 4, * and Agnieszka Surowiak 4
1 Faculty of Mining, Petroleum and Geophysics, Shahrood University of Technology, Shahrood 3619995161,
Iran; aazizi@shahroodut.ac.ir (A.A.); masdaryan_engineer@yahoo.com (M.M.)
2 Department of Processing, Helmholtz Institute Freiberg for Resource Technology,
Helmholtz-Zentrum Dresden-Rossendorf, Chemnitzer Straße 40, 09599 Freiberg, Germany;
a.hassanzadeh@gmx.de
3 Department of Control and Modeling of Mineral Processing Systems, Institute of Mineral Processing,
ACECR at Tarbiat Modares, Tehran 1411883684, Iran; bahri.zahra@gmail.com
4 Department of Environmental Engineering, Faculty of Mining and Geoengineering,
AGH University of Science and Technology, al. Mickiewicza 30, 30-059 Krakow, Poland; asur@agh.edu.pl
* Correspondence: tniedoba@agh.edu.pl; Tel.: +48-126-172-056

Received: 6 July 2020; Accepted: 23 July 2020; Published: 26 July 2020 

Abstract: The dominant challenge of current copper beneficiation plants is the low recoverability of
oxide copper-bearing minerals associated with sulfide type ones. Furthermore, applying commonly
used conventional methodologies does not allow the interactional effects of critical parameters in the
flotation processes to be investigated, which is mostly overlooked in the literature. To tackle this issue,
the present paper aimed at characterizing the behavior of five key effective factors and their interactions
in a sulfidized copper ore. In this context, dosage of collector (sodium di-ethydithiophosphate,
60–100 g/t), depressant (sodium silicate, 80–120 g/t) and frother (methyl isobutyl carbinol (MIBC),
6–10 g/t), pulp pH (7–11) and agitation rate (900–1300 rpm) were examined and statistically analyzed
using response surface methodology. Flotation experiments were conducted in a Denver type agitated
flotation cell at the rougher stage. The experimental results showed that increasing the pH (from 8
to 10) at low agitation rate (1000 rpm) enhanced the recovery from 80.36% to 85.22%, while at high
agitation rate (1200 rpm), a slight declination occurred in the recovery. Meanwhile, increasing the
collector dosage at a lower frother value (7 g/t), caused a reduction of about 4.44% in copper recovery
owing to the interactions between factors, whereas at a higher frother level (9 g/t), the recovery was
almost unchanged. The optimization process was also performed using the goal function approach,
and maximum copper recovery of 92.75% was obtained using ~70 g/t collector, 110 g/t depressant,
7 g/t frother, pulp pH of 10 and 1000 rpm agitation rate.

Keywords: sulfidized copper ore; flotation; recovery; optimization; interaction effects

1. Introduction
Copper is one of the most applicable elements in the metallurgical industries with a production of
10 Mt annually [1]. Treatment operations for copper extraction mostly focus on sulfide or oxide ores
owing to the well-established metallurgical processes. With declining copper cut-off grades in sulfide-
or oxide-type ore resources, the mixed sulfide-oxide deposits can be considered as an alternative and
important source. However, their treatment from the mineral processing perspective is challenging
with respect to the complexity of process mineralogy. It is estimated that about 80% of copper is
produced from primary ores by the froth flotation, smelting, and refining and the rest of 20% is achieved
by hydrometallurgical methods [2].

Minerals 2020, 10, 660; doi:10.3390/min10080660 www.mdpi.com/journal/minerals


Minerals 2020, 10, 660 2 of 19

Flotation is one of the most widely used techniques to separate valuable minerals from gangue
minerals in the mineral processing, because of its ability to treat low-grade and complex raw materials
in the fine particle size ranges [3–7]. Froth flotation takes place in the presence of three different phases,
including solid (particles), liquid (water) and gas (air bubbles). The physical and chemical properties
of each phase play a fundamental role in the efficiency of the flotation process. This process for the
mixed sulfide-oxide ores is traditionally carried out with two distinctive treatments including direct
flotation by fatty acids and sulfidizing by a sulfidization agent such as Na2 S, NaHS, and (NH4 )2 S [8–10].
A principal issue involved in the flotation of mixed sulfide-oxide ores and the copper industry is
the low recovery of oxide section. This is related to the surface properties of copper oxide minerals,
which cannot respond well to traditional sulfide copper collectors [11].
Moreover, flotation operations depend on many factors including the minerals’ nature and
structure (mineralogy, morphology and particle size), water chemistry, bubble size and velocity, particle
residence time, stator and rotor configurations, hydrodynamic properties, pulp potential and pH,
pulp density, air flowrate, and reagent types and dosages [3,6,12–20]. An understanding of the influence
of these factors is essential to improve the flotation circuit performance. Technically, grade, recovery
and selectivity index are commonly used as the most critical metallurgical parameters for this purpose.
An increase of 1−2% in these parameters is economically very significant [17] and plays a remarkable
role in the desirability of the flotation plants. In fact, an inefficient operation can cause an incorrect
balance between grade and recovery, which leads to adverse shifts in the grade-recovery curve [21,22].
Dhar et al. [23] studied the application of xanthate, dithiophosphate, thionocarbamate collectors,
and their mixture for flotation of copper sulfides from Nussir ore. They found that xanthate
and thionocarbamate with the ratio of 1 to 3 had the highest grade (29%) and copper recovery
(98%). Hangone et al. [24] showed that the maximum copper recoveries could be obtained with the
di-ethyl-dithiophosphate compared to that obtained with other thiol collectors at the same equivalent
dosage. Asadi et al. [25] beneficiated a sequential sulfidization-flotation technique for copper oxide
flotation taken from the Chahmora mine situated in Semnan province, Iran. They employed response
surface methodology for evaluating and optimizing process performance. Their results indicated that
a three-stage sulfidization was suitable for copper flotation to acquire a rougher concentrate with the
grade and recovery of 9.28% and 81.15% under optimal conditions after 5 min sulfidization conditioning.
Tijsseling et al. [26] examined the flotation behavior of mixed oxide with sulfide copper-cobalt minerals
using different collectors. Their findings demonstrated that a dithiophosphate collector could recover
over 90% of the copper sulfides, and 70% of the copper oxide minerals. Asghari et al. [27] investigated
the improvement of copper recovery in an industrial flotation circuit (Sarcheshmeh copper mine),
and stated that about 95% of overall copper lost in final tailing was in the rougher circuit while
the rougher flotation recovery of mixed copper ore was 8% less than the sulfide ore. Additionally,
the increase of collector and NaHS dosages along with the size reduction of coarse particles (>74 µm)
of rougher tailing had a positive influence on the recovery. Following this, Ghodrati et al. [28] applied
response surface methodology for the optimization of chemical reagents used in Miduk porphyry
copper flotation. They reported that nearly 91.4% copper with the grade of 8.13% was recovered
under optimal conditions including 5 g/t Z11, 34 g/t C-4132, 5 g/t A3477, 15 g/t AF65, and 8 g/t MIBC.
Ávila-Márquez et al. [29] examined the flotation of a copper sulfide under acidic conditions using
a xanthogen formate compound as a collector. They argued that the highest copper recovery was
achieved about 97% at an initial collector concentration of 20 mg L−1 , a flotation time of 3 min, and pH of
2. The surface oxidation of mineral in copper sulfide flotation was investigated via Moimane et al. [30].
The investigations showed that the surface oxidation of copper sulfide minerals had some detrimental
effects on their flotation performance. Also, it was identified that chalcocite flotation was much more
sensitive to surface oxidation in comparison to chalcopyrite.
As considered in the literature, many research works have been performed on the optimization
and evaluation of the main factors influencing the flotation process [6,17,25–28,31], but the information
on the interaction between influential factors is limited. Meanwhile, each flotation operation has its
Minerals 2020, 10, 660 3 of 19

own specifications and mineralogical properties based on its particular restrictions to achieve the
highest recovery and, therefore, requires more detailed investigations. The present research is an
attempt to fill this gap. Moreover, many experiments must be carried out to determine the optimal
values of influential operating factors on the flotation performance, which is impossible due to the
massive number of flotation condition sets. Hence, in this research work, the main and first-order
interaction effects of the influential parameters on flotation behavior are characterized using response
surface methodology combined with central composite design as an experimental design technique.

2. Theoretical Background
The conventional optimization technique of a multivariable system follows one factor at a time and
its major disadvantage is disregarding interactive effects among the factors. Accordingly, it does not
imply the complete effects of various factors on the process [32–38]. Additionally, this approach requires
more data to determine optimal level, which is a time-consuming process and undesirable [39]. Thus,
design of experiment (DOE) can be used for optimizing such multivariable systems as its successful
application has been proven by many researchers in the mineral processing industry [17,40–42].
Response surface modeling (RSM) is one of the efficient techniques of DOE, which involves a
combination of mathematical and statistical methods based on the multivariate non-linear model
in which all factors are varied over a set of experimental runs. This technique is an invaluable,
cost-effective and practical tool for analysis of processes and quantification of the relationship between
the controllable inputs and the response surfaces (outputs) obtained even in the presence of complex
interactions [32,33,43–45].
Generally, response surface methodology process involves five major steps containing (i) designing
of a series of experiments, (ii) developing a mathematical model with the best fittings for a functional
relationship between input factors and outputs, (iii) finding the optimal set of experimental factors
that produce a maximum or minimum value of the response(s), (iv) predicting the response(s) and
checking the adequacy of the model within the setup of the experiments, and (v) representing the direct
and interaction effects of factors [42–45]. This technique includes different designs such as central
composite design (CCD), Box–Behnken design (BBD), three-level factorial, and Doehlert, which CCD is
one of the most popular designs used in the process optimization [32,33,45]. It is known that the CCDs
with fractional factorial points are the best option to build statistical models in terms of the number of
required experiments and quality of the data obtained [46].
In a CCD, the number of experiments (N) is determined according to Equation (1):

N = 2(n−p) + 2n + nc (1)

where n is the number of factors and p is fraction of the number of factors (p = 0 is for full factorial
design), nc is number of central runs to estimate experimental error and 2n represents axial runs.
In addition, all factors are studied in five levels (−α, −1, 0, +1, +α), where the value of α is the star
(axial) point and is measured by the following formula [45,47,48]:
1
α = (2(n−p) ) 4 (2)

Two significant models, including first-order model (Equation (3)) and second-order polynomial
model (Equation (4)) are typically applied in RSM methodology.

k
X
Y = β0 + βi xi + ε (3)
i=1

k
X k
X k
X
Y = β0 + βi xi + βii x2i + βi j xi x j + ε (4)
i=1 i=1 1≤i≤ j
Minerals 2020, 10, 660 4 of 19

in which Y exhibits the predicted response; k depicts the number of factors; β0 displays a constant term;
βi denotes the linear coefficients; xi represents the independent variables or factors; βii depicts the
quadratic coefficients; βij indicates the interaction coefficients; and ε is the error.
In this study, the experimental data are analyzed statistically using Design Expert software
(Demo version 7.0.0, from Stat-Ease Inc., Minneapolis, MN, USA) and factors are studied with their
codified values for uniform comparison according to the following equation:

Xi − X0
xi = (5)
∆X
where xi implies the dimensionless coded value of the ith factor, Xi denotes the actual value of the
factor, X0 represents the value of Xi at the center point and ∆X is the step change value [41].

3. Experimental Procedure

3.1. Material and Grinding and Sulfidization Processes


The initial mixed sulfide-oxide copper bulk sample (nearly 100 kg) was provided from Abbas
Abad copper mine, located in the northeast of Semnan Province, Iran. The samples were prepared
after two stages of laboratory comminution including crushing in a jaw crusher and a roll crusher
followed by grinding by a standard Bond ball mill. Firstly, the given sample was crushed using
a jaw crusher (Fritsch 01.703, Fritsch, Idar-Oberstein, Germany) to reach a particle size finer than
10 mm. After that, samples were crushed twice by a roll crusher to achieve particles less than
4 mm in diameter. The samples were homogenized and subsequently split by means of cones
and riffles. Ultimately, the representative samples were chemically analyzed via X-ray fluorescence
spectrometer (XRF Shimadzu-1800, Shimadzu Corporation, Kyoto, Japan) and X-ray diffraction (XRD,
Unisantis XMD300, Unisantis, Georgmarienshűtte, Germany) as shown in Tables 1 and 2, respectively.

Table 1. Chemical analysis of the representative sample obtained from Abbas Abad copper mine
(wt.%).

Composition SiO2 CaO Fe2 O3 K2 O MnO TiO2 BaO Al2 O3


Content 52.10 6.90 6.34 4.52 0.21 0.71 0.09 17.64
Composition SO3 P2 O5 MgO Sr Na2 O Cu * L.O.I **
Content 0.19 0.61 2.71 0.11 2.66 0.42 4.79
* It composed of 0.22 (wt.%) copper sulfide and 0.2 (wt.%) copper oxide. ** L.O.I. is loss on ignition.

Table 2. The X-ray diffraction (XRD) analysis results obtained from Abbas Abad copper ore
sample (wt.%).

Compound Chemical Composition S-Q System


Anorthite, sodian, ordered (Ca,Na)(Al,Si)2 Si2 O8 16.4 Triclinic
Albite, calcian, ordered (Na,Ca)Al(Si,Al)3 O8 13.2 Triclinic
Microcline, intermediate KAlSi3 O8 11.7 Triclinic
Sanidine, disordered K(Si3 Al)O8 10.5 Monoclinic
Orthoclase KAlSi3 O8 9.4 Monoclinic
Magnesioferrite, high, syn MgFe2 O4 7.0 Cubic
Diopside, aluminian Ca(Mg,Fe,Al)(Si,Al)2 O6 5.8 Monoclinic
Calcite, syn CaCO3 5.3 Hexagonal
Magnetite FeO·Fe2 O3 4.4 Cubic
Montmorillonite-14A Na0.3(Al,Mg)2 Si4 O10 (OH)2 ·xH2 O 3.1
Bytownite, low 0.23NaAlSi2 O8 ·0.77CaAl2 Si2 O8 2.5 Triclinic
Andesine, low 0.62NaAlSi2 O8 ·0.38CaAl2 Si2 O8 2.4 Triclinic
Chlorite (Mg,Fe)5 (Al,Si)5 O10 (OH)8 2.3 Monoclinic
Biotite-2 ITM1 RG KMg3 (Si3 Al)O10 (OH)2 2.1 Monoclinic
Hematite Fe2 O3 2.0 Hexagonal
Kaolinite 1T Al2 Si2 O5 (OH)4 1.9 Triclinic
Minerals 2020, 10, 660 5 of 19

Table 1 presents the chemical characterization of the representative samples using XRF, and it
confirms the presence of 0.42 (wt.%) copper in the sample studied. Copper minerals of the ore exist in
two forms of both oxide (malachite) and sulfide (chalcocite and covellite) minerals. More detailed
information in this regard can be found elsewhere [5]. Table 2 demonstrates the XRD analysis of the
representative sample. The results demonstrate that the main gangue minerals are silicates and also
altered minerals of aluminosilicate, montmorillonite and kaolinite are found in the ore.
Then, 1100 g representative samples were ground by a laboratory Bond ball mill (12 × 5 in, Denver,
England) at an optimal time of 22 min [5] with a solid percentage of 70% w/w to reach 80% passing
∼80 µm. This particle range is identified as the most favorable size for copper flotation in several
works conducted by the authors [49–52]. The sieve analysis of mill product (flotation feed) is presented
in Table 3. It should be mentioned that samples studied are in the oxide and sulfide forms (50% copper
sulfide and 50% copper oxide) and their simultaneous flotation is difficult due to different behavior
of oxide and sulfide minerals. Therefore, the mechanochemical sulfidization process was applied to
sulfidize the oxide copper in the grinding stage. Milling was carried out on the representative sample
by adding the sulfur at a value of 0.5% based on ball mill feed weight. We presented this novel process
of sulfidization in our previous work, which detailed information was provided there [5].

Table 3. Sieve analysis of flotation feed sample studied.

Mass Retained on % Weight Retained on Cumulative % Cumulative %


Sieve Size (µm)
Each Sieve (g) Each Sieve Weight Retained Weight Passing
2000 4.38 1.41 1.41 98.59
841 1.85 0.59 2.00 98.00
595 0.71 0.23 2.23 97.77
400 0.50 0.16 2.39 97.61
297 1.05 0.34 2.73 97.27
210 3.50 1.12 3.85 96.15
149 7.20 2.31 6.16 93.84
105 18.60 5.97 13.13 87.87
74 30.30 9.72 21.85 78.15
53 28.50 9.14 30.99 69.01
37 45.10 14.97 45.46 54.54
−37 170.03 54.55 100.00 0.00
Sum 311.72 100.00

3.2. Flotation Experiments


Rougher flotation experiments were carried out using a 3 L Denver flotation machine with a solid
percentage of about 30% w/w at the desired impeller speed (900–1300 rpm). Lime (Ca(OH)2 ) was
applied to modify the pH at a targeted level (7–11) during the flotation process. It should be noted that
all experiments were carried out using tap water. First, sodium silicate was utilized as a depressant
agent of silicate gangue minerals at predetermined values of 80–120 g/t and after 5 min sodium
di-ethyldithiophosphate (DTP, Na3 PS2 O2 ) was added as a collector at desired values (60–100 g/t). Then,
after 3 min conditioning, MIBC (methyl isobutyl carbinol, (CH3 )2 CHCH2 CH(OH)CH3 ) as broadly
used frother was added into the cell at required values (6–10 g/t). After 1 min, flotation cell was aerated,
and froth was collected for 13 min with scraping interval of every 20 s. After each test, the tailings and
concentrates were filtered, dried, weighed and assayed for copper content.

4. Results and Discussions

4.1. Design of Experiments


According to the RSM-CCD methodology, 29 sets of experiments (N = 2(5 − 1) + 2 × 5 + 3 = 29) with
a suitable combination of the amounts of collector, depressant, frother, pH and agitation speed were
designed and conducted. The factors and their levels in coded and actual units are presented in Table 4.
It needs to be pointed out that the length of the axial (star) points was chosen based on Equation (2);
Minerals 2020, 10, 660 6 of 19

(α = (2(5 − 1) )1/4 = 2). Table 5 also shows the design matrix of the variables and the corresponding
responses for each experiment.

Table 4. The main factors influencing the copper flotation process and their levels.

Coded Levels
Factors Symbol −2 −1 0 +1 +2
Actual Levels
Collector dosage (g/t) A 60 70 80 90 100
Depressant dosage (g/t) B 80 90 100 110 120
Frother dosage (g/t) C 6 7 8 9 10
pH D 7 8 9 10 11
Agitation speed (rpm) E 900 1000 1100 1200 1300

Table 5. Central composite design (CCD) matrix and experimental (measured) values of copper recovery.

Run A B C D E Cu Recovery (%)


1 90 110 7 10 1000 86.90
2 80 100 8 11 1100 86.51
3 80 100 10 9 1100 82.14
4 80 100 8 9 1100 82.32
5 70 110 9 10 1000 82.19
6 90 110 9 10 1200 83.53
7 60 100 8 9 1100 79.60
8 80 100 6 9 1100 86.27
9 70 90 7 8 1200 86.55
10 70 90 9 8 1000 79.17
11 90 90 7 8 1000 83.47
12 90 90 9 8 1200 85.74
13 70 90 9 10 1200 80.9
14 70 110 7 10 1200 87.64
15 80 100 8 9 1100 82.28
16 90 110 7 8 1200 81.00
17 70 90 7 10 1000 92.4
18 90 90 7 10 1200 83.71
19 70 110 7 8 1000 87.21
20 80 100 8 9 900 84.64
21 70 110 9 8 1200 88.34
22 90 110 9 8 1000 78.35
23 90 90 9 10 1000 85.03
24 80 120 8 9 1100 84.32
25 100 100 8 9 1100 76.88
26 80 100 8 9 1300 84.98
27 80 100 8 7 1100 81.74
28 80 100 8 9 1100 78.98
29 80 80 8 9 1100 77.45

Since the objective was to investigate and optimize the influential operational factors on the
flotation of copper, the copper recovery, R (%) was considered as the process response and calculated
by Equation (6).
C c
R = 1 × 100 (6)
Ff
where F (g) and C1 (g) depict the dry weight of feed and concentrate and f (%) and c (%) represent the
% grade of feed and concentrate, respectively [12].
Minerals 2020, 10, 660 7 of 19

4.2. Model Construction for Cu Recovery


To understand the behavior of factors and process, it is essential to select an appropriate model.
Hence, various models were fitted to the flotation data and assessed by analysis of variance (ANOVA)
in terms of the quality and accuracy of the models fitted (Table 6).

Table 6. Statistical analysis of models for modeling rougher flotation of copper.

Source SS DF MS F-Value p-Value


Analysis of Variance (ANOVA) of Models
Mean versus Total 2.02 × 105 1 2.02 × 105
Linear versus Mean 94.81 5 18.96 1.61 0.1985
2FI versus Linear 136.74 10 13.67 1.32 0.3153
Quadratic versus 2FI 93.75 5 18.75 3.64 0.0517 Suggested
Cubic versus Quadratic 21.25 5 4.25 0.64 0.6925 Aliased
Residual 19.99 3 6.66
Total 2.02 × 105 29 6977.64
Analysis of Lack of Fit
Linear 264.38 21 12.59 3.43 0.2502
2FI 127.64 11 11.6 3.16 0.265
Quadratic 33.89 6 5.65 1.54 0.445 Suggested
Cubic 12.64 1 12.64 3.44 0.2048 Aliased
Pure Error 7.35 2 3.67
Statistical Analysis Summary of Models
Source Std. Dev. R-Square PRESS
Linear 3.44 0.2587 443.19
2FI 3.22 0.6317 1579.43
Quadratic 2.27 0.8875 873.73 Suggested
Cubic 2.58 0.9455 14,359.54 Aliased
SS—sum of squares; DF—degrees of freedom; MS—mean squares; PRESS—predicted residual error sum of squares.

The ANOVA was introduced by Fisher, who developed it to solve agricultural problems. It enables
an assessment of the influence of certain controlled factors on the outcome of an experiment. An analysis
of variance can be carried out according to two schemes. The first is to study the variance for the single
classification; the second scheme provides for multivariate data analysis. However, the fundamental
disadvantage of ANOVA is considered as the impact of only one variable dependent on a certain
phenomenon. Wilks met the problem with his multidimensional analysis of variance (MANOVA),
which allows a simultaneous examination of relations between two or more independent variables of
type and two or more dependent variables of quantitative type. Hence, both ANOVA and MANOVA
turn out to be useful in the study of grained mineral properties and their impact on the beneficiation
effects [53–56].
Table 6 shows that a quadratic model is more significant than other models. According to
ANOVA in Table 6, F-value for the model (3.64) and the lack of fit is 1.54 indicating that a quadratic
polynomial model with R2 of 0.8895 is adequate and capable in approximating the response surface
and representing the system under the given experimental domain. Therefore, a quadratic polynomial
model (Equation (4)) was fitted to experimental data. After removing the very insignificant interaction
terms (p-value > 0.1), the final equation representing the recovery of Cu was achieved as a function of
collector (A), depressant (B) and frother (C) dosages, pH (D) and agitation speed (E). Equation (7) is
the final formula developed based on coded factors:

R= 80.51 − 0.92 × A + 0.50 × B − 1.41 × C + 0.92 × D + 0.14 × E − 0.91 × A × B


+1.3 × A × C + 1.55 × C × E − 1.51 × D × E + 1.14 × C2 + 1.12 × D2 (7)
+1.29 × E2

From the analysis of Equation (7), because of the individual variables (A, B, C, D and E) and
assuming the remaining four mean values equal to zero, it occurs that the maximum values do
Minerals 2020, 10, 660 8 of 19

not exceed 86.83% (Rmax (A) = 81.43%; Rmax (B) = 81.51%; Rmax (C) = 82.25%; Rmax (D) = 86.83%;
Rmax (E) = 85.98%), while Ropt = 92.40%, which is the confirmation that the interactions between
individual variables influence the result significantly.

4.3. Model Validation and Statistical Evaluation


After modeling, the data were statistically analyzed at 95% confidence level to evaluate the
significance of the model and determine the rank of important factors. ANOVA was used for this
purpose, the results of which are presented in Table 7. As considered in Table 7, R2 is 0.8097,
which indicates that the model can explain 80.97% of the variability in Cu recovery. The predictive
copper recoveries by the model (Equation (7)) were plotted against the actual values (Figure 1a).
Additionally, the values of the conformity factor η2 were listed in Table 8, which expresses the
percentage of variance explained by the effect of individual and combined factors.

Table 7. ANOVA for response surface modeling (RSM) quadratic model.

Source SS DF MS F-Value p-Value


Model 296.79 12 24.73 5.67 0.0009 Significant
A-Collector (g/t) 20.37 1 20.37 4.67 0.0461
B-Depressant (g/t) 5.93 1 5.93 1.36 0.2606
C-Frother (g/t) 47.86 1 47.86 10.98 0.0044
D-pH 20.19 1 20.19 4.63 0.0470
E-Agitation rate (rpm) 0.47 1 0.47 0.11 0.7461
AB 13.2 1 13.2 3.03 0.1011
AC 26.96 1 26.96 6.19 0.0243
CE 38.6 1 38.6 8.85 0.0089
DE 36.51 1 36.51 8.38 0.0106
C2 35.12 1 35.12 8.06 0.0119
D2 33.9 1 33.9 7.78 0.0131
E2 45.03 1 45.03 10.33 0.0054
Residual 69.75 16 4.36
Lack of Fit 62.4 14 4.46 1.21 0.5413 Not significant
Pure Error 7.35 2 3.67
Cor Total 366.54 28
Std. Dev. 2.09
R-Squared 0.8097
Adequate Precision 9.57
PRESS 218.84

Predicted vs. Actual e Normal Plot of Residuals


93.00
99

95
88.75 90
Normal % Probability

80
70
Predicted

50
84.50
30
20
10
5
80.25

(a) 1
(b)
76.00
-1.58 -0.78 0.03 0.83 1.63
76.75 80.67 84.58 88.49 92.40
Internally Studentized Residuals

Actual

Figure 1. The plot of actual (measured) versus predicted values for copper recovery in (a) rougher
flotation process and (b) plot of normal probability versus Studentized residual.

‐ ‐




Minerals 2020, 10, 660 9 of 19

Table 8. The proportion of total variance explained by the certain experimental effect.

η2 model η2 C η2 E2 η2 CE η2 DE η2 C2 η2 D2 η2 AC η2 A η2 D Σ (%)
80.97 12.78 12.03 10.31 9.75 9.38 9.06 7.20 5.44 5.39 81.34

Adequate precision comprises the predicted value at the design points, the average prediction
error and a ratio greater than 4 is desirable [6,38]. In this study, it was found to be a ratio of 9.57 (Table 7),
which shows the model’s ability. The adequacy of the predictive model generated for the copper
recovery is further demonstrated from the residual normal probability plot in Figure 1b. The data
presented in Figure 1b closely follows the straight line showing that there is no requirement of actual
data transformations. It can be concluded that the deviation between the predicted and experimental
values are distributed normally and the suggested model is well-fitted to data. This further supports
the validity of the second-order model to predict copper recovery.
In addition to the model adequacy, ANOVA was employed to determine the relative significance
and the influential degree of factors. The essential parameters were ranked based on the p-, F-values
and conformity factor η2 . It can be considered from Tables 7 and 8 that the significance degree of the
important terms is in the order of frother > quadratic effect of agitation rate > interactive effect between
frother and agitation speed > interaction effect between pH and agitation speed > quadratic effect
of frother > quadratic effect of pH > interactive effect between collector and frother > linear effect of
collector > linear effect of pH.

4.4. Influence of Factors and Their Interactions on Recovery


The 3D response surface graphs were depicted to identify the individual, interaction effects and
generally combined impacts of factors studied on the copper recovery (Figures 2–5).
Figure 2 indicates the 3D response surface (Figure 2b), contour plots (Figure 2a) and the interactive
impact (Figure 2c) of collector and depressant on the copper flotation when other experimental factors
were constant at the central values (pH = 9, frother dosage = 8 g/t and agitation rate = 1100 rpm).
It can be seen in Figure 2 that, at low levels of collector dosage (70 g/t), increasing the value of
depressant (sodium silicate) from 90 to 110 g/t results in an improvement in the recovery of copper
(from 80.03% to 82.84%). This might be related to the fact that the presence of sodium silicate in
flotation pulp leads to the increase in the load on the particle surface, favorable dispersion of fine
particles and consequently satisfactory results of flotation performance. Moreover, a higher dosage of
sodium silicate helps to upgrade the flotation performance by the depression of silicate minerals [25,57].
According to Park and Jeon [58], sodium silicate is habitually used for a dispersant/depressant. Based on
the chemical reactions (Equations (8)–(10)), three anionic species are formed in the pulp, viz. OH− ,
2− 2−
HSiO− − −
3 , and SiO3 . The OH ion enhances the pH of pulp, while HSiO3 and SiO3 are the main
depressing/dispersing species of silicates. A stable dispersion of the particles in the pulp is attributed to
repulsion between anionic species and minerals. This phenomenon desirably prevents a slime coating
on the mineral surfaces. Consequently, it promotes the adsorption rate of collectors and subsequently,
copper floatability.
Na2 SiO3 + 2H2 O ↔ H2 SiO3 + 2Na+ + 2OH− (8)
+
H2 SiO3 ↔ HSiO−
3 +H (9)
2− +
HSiO−
3 ↔ SiO3 + H (10)

Furthermore, the increase in DTP collector dosage in the range studied (70–90 g/t) leads to a
reduction in the recovery (Figures 2 and 3), because of the micellization phenomenon for the additional
values of the collector in the range investigated [59,60].
Minerals 2020, 10, 660 10 of 19

Recovery (%)
90.00

79.7893
82.9

85.00
81.95

Recovery (%)
Collector (g/t)

81
80.399
80.00

80.05
81.0088
79.1
75.00
81.6186

82.2284 90.00 110.00


85.00 105.00
70.00
80.00 100.00
90.00 95.00 100.00 105.00 110.00
Collector (g/t) 75.00 95.00 Depressant (g/t)
70.00 90.00
Depressant (g/t)

(a) (b)

A: Collector (g/t)
93

A- 70.000
A+ 90.000
88.75
B: Depressant (g/t)
Recovery (%)

A: Collector (g/t)

Actual Factors
84.5 C: Frother (g/t) = 8.00
D: pH = 9.00
E: Agitation rate (rpm) = 1100.00

80.25

76

90.00 95.00 100.00 105.00 110.00

B: Depressant (g/t)

(c)

Figure 2. (a) Contour and (b) 3D response surface plots showing the combined effect of collector and
depressant and (c) their interaction effect on the copper recovery.

According to Dhar et al. [61], the mechanism of copper mineral flotation with dithiophosphates
is in conjunction with chemical adsorption by the formation of cuprous DTP in neutral and slightly
alkaline conditions. Also, Roy et al. [62] reported that the adsorption of this collector type occurs
very quickly on the surface of sulfide minerals and at the first few seconds after the addition of
the collector. Meanwhile, for the frother, a similar trend with the collector is observable (Figure 3).
Although there is a linear relationship between Cu recoveries with collector dosage, it is quadratic for
the frother. Figure 3a–c demonstrates that at low levels of frother (7 g/t), the increment in the amount
of collector causes attenuation in copper recovery (from 85.29% to 80.85%) and at low values, it was
almost unchanged. Furthermore, an increase in the collector dosage (from 70 to 90 g/t) leads to a slight
increase in copper recovery. According to Figure 3a–c, the maximum copper recovery is obtained at
the collector and frother dosages of 70 g/t and 7 g/t, respectively. It can also be seen that changing the
collector consumption results in only a minor influence on the flotation of copper in the range tested;
Minerals 2020, 10, 660 11 of 19

however, it was statistically significant. This effect can be described by the micellization phenomenon
in the range tested [59].

Recovery (%)
90.00

85.4

85.00
83.925

Recovery (%)
Collector (g/t)

80.5373
82.45
80.00 81.487
80.975
82.4366
79.5
75.00 83.3863

84.3359 90.00 9.00


85.00 8.50
70.00
80.00 8.00
7.00 7.50 8.00 8.50 9.00
Collector (g/t) 75.00 7.50 Frother (g/t)
70.00 7.00
Frother (g/t)

(a) (b)

C: Frother (g/t)
93
C- 7.000
C+ 9.000
A: Collector (g/t)
88.75 C: Frother (g/t)
Actual Factors
B: Depressant (g/t) = 100.00
Recovery (%)

D: pH = 9.00
E: Agitation rate (rpm) = 1100.00
84.5

80.25

76

70.00 75.00 80.00 85.00 90.00

A: Collector (g/t)
(c)

Figure 3. (a) Contour and (b) 3D response surface plots showing the combined effect of collector and
frother and (c) their interaction effect on the copper recovery.

Figures 4 and 5 exhibit the response surface graphs of the copper recovery as a function of agitation
rate with frother and pH. It is seen that the recovery is very sensitive and effective to change in the
agitation rate, frother dosage and pulp pH, and there is a quadratic relationship between these factors
with copper recovery.
According to Fornasiero and Filippov [63] as well as Hassanzadet et al. [64] a high impeller speed in
the flotation cell is essential for maximum particle–bubble collision frequency, but an intensive agitation
is detrimental as‐ the coarser particles detach from the bubbles and recovery reduces. According to
Figure 4, at high levels of frother (9 g/t), the increase of the agitation rate from 1000 to 1200 rpm causes
an increase in copper recovery from 79.84% to 83.22%. Also, the findings in Figure 5 indicate that at
low levels of agitation speed (1000 rpm), copper recovery enhances from 80.36% to 85.22% when pH
exceeds from 8 to 10. Increasing the copper recovery into the concentrate at the highest pH-value is
related to the stability of froth and the greater ability to carry particles by bubbles. The optimal pH
was found to be 10.
Minerals 2020, 10, 660 12 of 19

Recovery (%)
9.00

85.9
8.50
84.325

Recovery (%)
Frother (g/t)

82.75
8.00
80.7019
81.175

79.6
7.50
81.716
82.7301
83.7443 9.00 1200.00
84.7584 8.50 1150.00
82.7301
7.00 8.00 1100.00
1000.00 1050.00 1100.00 1150.00 1200.00 Frother (g/t) 7.50 1050.00Agitation (rpm)
7.00 1000.00
Agitation (rpm)

(a) (b)

C: Frother (g/t)
93
C- 7.000
C+ 9.000
E: Agitation rate (rpm)
C: Frother (g/t)
88.75 Actual Factors
A: Collector (g/t) = 80.00
Recovery (%)

B: Depressant (g/t) = 100.00


D: pH = 9.00
84.5

80.25

76

1000.00 1050.00 1100.00 1150.00 1200.00

E: Agitation rate (rpm)


(c)

Figure 4. (a) Contour and (b) 3D response surface plots showing the combined effect of frother and
agitation rate and (c) their interaction effect on the copper recovery.

According to Somasundaran and Wang [65], Liu et al. [66] and Gu et al. [67], the presence of sulfur
ions is a function of pH. It can be concluded that H2 S, HS– and S2– form species are predominant
at low (pH < 7), medium (pH 7–12) and high pH > 12 pH range, respectively. Therefore, it can be
concluded that the optimal pH of flotation is located in the range where HS− is predominant species.
HS− is easily absorbed on the surface of oxide minerals and produces a sulfidized surface film (such as
CuS) and subsequently activates the oxides flotation. It is also mentioned that generally, the copper
flotation takes place in the area on the Eh-pH diagram where the hydrophobic species of CuEX(s) is
formed. In contrast, the depression of copper occurs in the stability area of Cu(OH)2 (s) and (EX)2 (aq)
compounds. This behavior indicates that the reduction of copper recovery in the values of pH >10 may
be owing to the formation of Cu(OH)2 (s) instead of CuEX(s) [25]. In this context, Fairthorne et al. [68]
demonstrated that reducing recovery at high pH may be due to the generation of fresh spots on mineral
surfaces during grinding, and the metal hydroxide precipitation. Additionally, Grano et al. [69,70]
expressed that surface of sulfide mineral (e.g., chalcopyrite) is significantly influenced by pulp pH
and the formation of iron oxyhydroxides on the surface of minerals at alkaline pHs prevents the
adsorption of DTP and depresses the flotation. Additionally, these behaviors can be further described
in Figures 4c and 5c, which show interaction effects between agitation speed with frother dosage and
pH, respectively.
Minerals 2020, 10, 660 13 of 19

Recovery (%)
10.00
83.524
81.833
82.6785
85.4
9.50
80.9875 84.075

Recovery (%)
82.75
pH

9.00
81.425

80.1
8.50

10.00 1200.00
82.6785 9.50 1150.00
9.00 1100.00
8.00
pH 8.50 1050.00Agitation (rpm)
1000.00 1050.00 1100.00 1150.00 1200.00
8.00 1000.00

Agitation (rpm)

(a) (b)

E: Agitation rate (rpm)


93

E- 1000.000
E+ 1200.000
88.75 D: pH
E: Agitation rate (rpm)
Recovery (%)

Actual Factors
A: Collector (g/t) = 80.00
84.5 B: Depressant (g/t) = 100.00
C: Frother (g/t) = 8.00

80.25

76

8.00 8.50 9.00 9.50 10.00

D: pH

(c)

Figure 5. (a) Contour plot and (b) 3D response showing the combined effect of pH and agitation speed
and (c) their interaction effect on the copper recovery.

4.5. Process Optimization


The flotation process parameters were optimized using the DX software package and goal function
approach to maximize the recovery of copper within the experimental range, the results of which are
displayed in Figure 6. Figure 6 demonstrates the optimal conditions suggested by the software. It is
found that the maximum copper recovery can be obtained around 92.75% with a desirable amount of
100% (Desirability = 1). The optimal operating conditions suggested by DOE software were: collector
dosage of ~70 g/t, depressant amount of ~110 g/t, frother dosage of 7 g/t, pulp pH of ~10 and agitation
rate of ~1000 rpm. Two confirmation experiments were also performed to validate the model suggested
under optimal condition. The recovery average of two flotation tests was determined to be 90.16%,
which indicates the high accuracy of the model.
Minerals 2020, 10, 660 14 of 19

Figure 6. The optimum values of factors to achieve the maximum Cu recovery.

Figure 7a,b imply the relative significance of factors and the trend of factors for movement towards
the optimal point, respectively. Perturbation plots help to compare the impact of all the factors at a
particular point in the design space. A steep slope or curvature in a factor exhibits that the response
variable (copper recovery) is sensitive to that factor. It is observed from Figure 7a the frother (C) and
collector (A) dosages and the value of pH are the most critical factors on the recovery of copper. It can
also be seen in Figure 7b that changing the value of factors toward the optimal point reduced the
recovery of copper. The copper recovery increased with an increase in the depressant (B) and pH (D)
dosages and also a decrease in the amount of the frother (C) and collector (A) amounts and agitation
rate (E). In addition, it can be considered that the recovery of copper was a linear function of A and B
and a quadratic function of C, D and E in the range studied. This behavior was already discussed in
previous sections from a technical point of view.

Perturbation Perturbation
85 93
B A
D CE
B
C
83 88.75
D D
E
Recovery (%)
Recovery (%)

E A
AE
B C
81 D 84.5

B C
A
79 80.25

77 76

-1.000 -0.500 0.000 0.500 1.000 -1.999 -0.999 0.000 1.000 1.999

Deviation from Reference Point (Coded Units) Deviation from Reference Point (Coded Units)

(a) (b)

Figure 7. Perturbation plot showing trend of factors on the recovery of copper before (a) and after (b)
optimization process.
Minerals 2020, 10, 660 15 of 19

5. Conclusions
The rougher flotation behavior of a sulfidized copper ore was studied by a series of batch flotation
experiments. Response surface modeling based on central composite design was applied to optimize
and evaluate the effects of collector, depressant and frother dosages, solution pH and agitation rate on
the copper recovery. ANOVA and 3D response surface plots were applied to investigate the behavior
of copper flotation. The major conclusions can be summarized as follows:

• A quadratic mathematical model (with R2 of 0.8097) was developed to predict the copper
recovery as a function of influential factors after validation and confirmation through several
strong methods.
• ANOVA indicated that the linear effects of collector, frother and pH, the interaction effects of
frother with collector and agitation rate and pH with agitation rate, the quadratic effects of frother,
pH and agitation were significant on the copper recovery. Also, the significance degree of factors
was obtained as frother dosage > agitation rate2 > frother × agitation speed > pH × agitation
speed > frother dosage2 > pH2 > collector × frother > collector dosage > pH.
• According to the 3D response surface graphs: (a) the recovery of copper improved about 2.81% by
increasing the amount of depressant (sodium silicate) at low levels of collector dosage (70 g/t);
(b) the copper recovery decreased by increasing the amount of dithiophosphate collector at high
levels of frother (about 4.44%) and conversely at low values, the increase of the collector value to
90 g/t had no significant effect on the recovery; (c) changing the collector dosage had limited effects
on the floatability of copper in the range tested and the recovery reduced with enhancing the
value of collector dosage; (d) at low values of agitation rate, raising the pH increased significantly
the copper recovery, while at high values of agitation, a slight reduction occurred in the recovery.
• Using an experimental design and goal function approach, the optimal levels of factors were
determined to be ~70 g/t for collector, 110 g/t for depressant, 7 g/t for frother, 10 for pulp pH and
1000 rpm for agitation rate with a prediction of 92.75% copper recovery. Under these conditions,
the desirability of the optimal condition was found to be 1. Confirmatory experiments were also
conducted to assess the model suggested under the optimal condition predicted, and the results
showed that the developed model was reliable and helpful for predicting the copper recovery.

Author Contributions: Conceptualization/methodology/investigation/writing—original draft preparation,


A.A.; methodology/investigation; M.M. and Z.B.; Conceptualization/writing—original draft preparation/
writing—review and editing; A.H.; T.N. and A.S. All authors have read and agreed to the published version of
the manuscript.
Funding: This research received no external funding.
Acknowledgments: The authors would like to appreciate the Abbas Abad copper mine for their support during
this research.
Conflicts of Interest: The authors declare no conflict of interest.

Nomenclature
RSM Response surface methodology
CCD Central composite design
DOE Design of experiment
BBD Box–Behnken design
DTP Di-ethyldithiophosphate
XRD X-ray diffraction
XRF X-ray fluorescence
MIBC Methyl isobutyl carbinol
R Recovery
C1 Dry weight of concentrate
Minerals 2020, 10, 660 16 of 19

c % grade of concentrate
F Dry weight of feed
f % grade of feed
η2 Conformity factor
ANOVA Analysis of variance
MANOVA Multidimensional analysis of variance
p-value Probability value (significance level)
R2 Coefficient of determination
SS Sum of squares
n and k Number of factors
p Fraction of the number of factors
nc Number of central runs
2n Axial runs
α Star (axial) point
Y Predicted response
β0 Constant term
βi linear coefficient
xi Independent variable
βii Quadratic coefficient
βij Interaction coefficient
ε Error
xi Dimensionless coded value
Xi Actual value
∆X Step change value
Cu Copper
MS Mean squares
DF Degrees of Freedom
Std. Dev. Standard deviation
PRESS Predicted residual error sum of squares

References
1. Ekman Nilsson, A.; Macias Aragonés, M.; Arroyo Torralvo, F.; Dunon, V.; Angel, H.; Komnitsas, K.;
Willquist, K. A Review of the carbon footprint of Cu and Zn production from primary and secondary sources.
Minerals 2017, 7, 168. [CrossRef]
2. Davenport, W.G.; King, M.; Schlesinger, M.; Biswas, A.K. Extractive Metallurgy of Copper; Elsevier: Oxford,
UK, 2002; pp. 307–325.
3. Rahman, R.M.; Ata, S.; Jameson, G.J. The effect of flotation variables on the recovery of different particle size
fractions in the froth and the pulp. Int. J. Miner. Process. 2012, 106, 70–77. [CrossRef]
4. Maldonado, M.; Araya, R.; Finch, J. An overview of optimizing strategies for flotation banks. Minerals 2012,
2, 258–271. [CrossRef]
5. Masdarian, M.; Azizi, A.; Bahri, Z. Mechanochemical sulfidization of a mixed oxide-sulphide copper ore
by co-grinding with sulfur and its effect on the flotation efficiency. Chin. J. Chem. Eng. 2020, 8, 743–748.
[CrossRef]
6. Azizi, A. Optimization of rougher flotation parameters of the Sarcheshmeh copper ore using a statistical
technique. J. Disper. Sci. Technol. 2015, 36, 1066–1072. [CrossRef]
7. Hassanzadeh, A.; Azizi, A.; Kouachi, S.; Karimi, M.; Celik, M.S. Estimation of flotation rate constant and
particle-bubble interactions considering key hydrodynamic parameters and their interrelations. Miner Eng.
2019, 141, 105836. [CrossRef]
8. Lee, K.; Archibald, D.; McLean, J.; Reuter, M.A. Flotation of mixed copper oxide and sulphide minerals with
xanthate and hydroxamate collectors. Miner. Eng. 2009, 22, 395–401. [CrossRef]
9. Yin, W.-Z.; Sun, Q.-Y.; Li, D.; Tang, Y.; Fu, Y.-F.; Yao, J. Mechanism and application on sulphidizing flotation
of copper oxide with combined collectors. Trans. Nonferrous Met. Soc. China. 2019, 29, 178–185. [CrossRef]
Minerals 2020, 10, 660 17 of 19

10. Conejeros, V.; Pérez, K.; Jeldres, R.I.; Castillo, J.; Hernández, P.; Toro, N. Novel treatment for mixed copper
ores: Leaching ammonia-Precipitation-Flotation (L.A.P.F.). Miner. Eng. 2020, 149, 106242. [CrossRef]
11. Zare, A.; Azizi, A.; Karamoozian, M. An investigation into the recovery of oxide copper from a complex
copper ore using sulphidisation technique and hydroxamate and potassium amyl xanthate collectors.
Geosys. Eng. 2020, 23, 43–50. [CrossRef]
12. Hoang, D.H.; Hassanzadeh, A.; Peuker, U.A.; Rudolph, M. Impact of flotation hydrodynamics on the
optimization of fine-grained carbonaceous sedimentary apatite ore beneficiation. Powder Technol. 2019, 345,
223–233.
13. Bulatovic, S.M. Handbook of Flotation Reagents Chemistry, Theory and Practice: Flotation of Sulfide Ores; Elsevier
Science & Technology: Amsterdam, The Netherlands, 2007.
14. Ucurum, M.; Bayat, O. Effects of operating variables on modified flotation parameters in the mineral
separation. Sep. Purif. Technol. 2007, 55, 173–181. [CrossRef]
15. Hassanzadeh, A.; Karakas, F. Recovery improvement of coarse particles by stage addition of reagents in
industrial copper flotation circuits. J. Disper. Sci. Technol. 2017, 38, 309–316. [CrossRef]
16. Zanin, M.; Wightman, E.; Grano, S.R.; Franzidis, J.P. Quantifying contributions to froth stability in porphyry
copper plants. Int. J. Miner. Process. 2009, 91, 19–27. [CrossRef]
17. Vazifeh, Y.; Jorjani, E.; Bagherian, A. Optimization of reagent dosages for copper flotation using statistical
technique. Trans. Nonferrous. Met. Soc. China. 2010, 20, 2371–2378. [CrossRef]
18. Azizi, A. Influence of collector dosage and pulp chemistry on copper flotation. Geosys. Eng. 2014, 17, 311–316.
[CrossRef]
19. Hassanzadeh, A.; Hassas, B.V.; Kouachi, S.; Brabcova, Z.; Çelik, M.S. Effect of bubble size and velocity in
chalcopyrite flotation. Colloids. Surf. A: Physicochem. Eng. Aspects. 2016, 498, 258–267. [CrossRef]
20. Agheli, S.; Hassanzadeh, A.; Vaziri Hassas, B.; Hasanzadeh, M. Effect of pyrite content of feed and
configuration of locked particles on rougher flotation of copper in low and high pyritic ore types. Int. J.
Mining. Sci. Technol. 2018, 28, 167–176. [CrossRef]
21. Singh, A.; Louw, J.J.; Hulbert, D.G. Flotation stabilization and optimization. J. South. Afr. Inst. Min. Metall.
2003, 103, 581–588.
22. Radmehr, V.; Shafaei, S.Z.; Noaparast, M.; Abdollahi, H. Optimizing flotation circuit recovery by effective
stage arrangements: A case study. Minerals 2018, 8, 417. [CrossRef]
23. Dhar, P.; Thornhill, M.; Rao Kota, H. Comparison of single and mixed reagent systems for flotation of copper
sulphides from Nussir ore. Miner. Eng. 2019, 142, 105930. [CrossRef]
24. Hangone, G.; Bradshaw, D.; Ekmekci, Z. Flotation of a copper sulphide ore from Okiep using thiol collectors
and their mixtures. J. South. Afr. Inst. Min. Metall. 2005, 105, 199–206.
25. Asadi, M.; Soltani, F.; Tavakoli Mohammadi, M.R.; Khodadadi Darban, A.; Abdollahy, M.
A successful operational initiative in copper oxide flotation: Sequential sulphidisation-flotation technique.
Physicochem. Probl. Miner. Process. 2019, 55, 356–369.
26. Tijsseling, L.T.; Dehaine, Q.; Rollinson, G.K.; Glass, H.J. Flotation of mixed oxide sulphide copper-cobalt
minerals using xanthate, dithiophosphate, thiocarbamate and blended collectors. Miner. Eng. 2019, 138,
246–256. [CrossRef]
27. Asghari, M.; Nakhaei, F.; VandGhorbany, O. Copper recovery improvement in an industrial flotation circuit:
A case study of Sarcheshmeh copper mine. Energ. Source. Part. A 2019, 41, 761–778. [CrossRef]
28. Ghodrati, S.; Nakhaei, F.; VandGhorbany, O.; Hekmati, M. Modeling and optimization of chemical reagents
to improve copper flotation performance using response surface methodology. Energ. Source. Part. A 2020,
42, 1633–1648. [CrossRef]
29. Ávila-Márquez, D.M.; Blanco-Flores, A.; Reyes-Domínguez, I.A.; Toledo-Jaldin, H.P.; Aguilar-Carrillo, J.;
Cruz-Gaona, R. Copper sulfide flotation under acidic conditions using a xanthogen formate compound as
collector: Adsorption studies and experimental design approach. Colloids Surf. A Physicochem. Eng. Asp.
2020, 585, 124032. [CrossRef]
30. Moimane, T.; Plackowski, C.; Peng, Y. The critical degree of mineral surface oxidation in copper sulphide
flotation. Miner. Eng. 2020, 145, 106075. [CrossRef]
31. Nakhaei, F.; Mosavi, M.R.; Sam, A.; Vaghei, Y. Recovery and grade accurate prediction of pilot plant flotation
column concentrate: Neural network and statistical techniques. Int. J. Miner. Process. 2012, 110–111, 140–154.
[CrossRef]
Minerals 2020, 10, 660 18 of 19

32. Montgomery, D.C. Design and Analysis of Experiments; John Wiley & Sons, Inc.: New York, NY, USA, 2001.
33. Myers, R.H.; Montgomery, D.C. Response Surface Methodology: Process and Product Optimization Using Designed
Experiments, 2nd ed.; John Wiley & Sons: New York, NY, USA, 2002.
34. Hinkelman, K.; Kempthorne, O. Design and Analysis of Experiments; John Wiley & Sons: Hoboken, NJ, USA,
2005.
35. Kincl, M.; Turk, S.; Vrecer, F. Application of experimental design methodology in development and
optimisation of drug release method. Int. J. Pharm. 2005, 291, 39–49. [CrossRef]
36. Simate, G.S.; Ndlovu, S.; Gericke, M. Bacterial leaching of nickel laterites using chemolithotrophic
microorganisms: Process optimisation using response surface methodology and central composite rotatable
design. Hydrometallurgy 2009, 98, 241–246. [CrossRef]
37. Niaki, R.; Abazarpoor, A.; Halali, M.; Maarefvand, M.; Ebrahimi, G. Application of Response Surface
Methodology and Central Composite Rotatable Design for Modeling and Optimization of Sulfuric and Nitric
Leaching of Spent Catalyst. Russ. J. Non-ferrous. Metals. 2015, 56, 155–164. [CrossRef]
38. Behera, S.K.; Meena, H.; Chakraborty, S.; Meikap, B.C. Application of response surface methodology (RSM)
for optimization of leaching parameters for ash reduction from low-grade coal. Int. J. Mining. Sci. Technol.
2018, 28, 621–629. [CrossRef]
39. Sahu, J.N.; Acharya, J.; Meikap, B.C. Response surface modeling and optimization of chromium (VI) removal
from aqueous solution using Tamarind wood activated carbon in batch process. J. Hazard. Mater. 2009, 172,
818–825. [CrossRef] [PubMed]
40. Rao, G.V.; Mohanty, S. Optimization of flotation parameters for enhancement of grade and recovery of
phosphate from low-grade dolomitic rock phosphate ore from Jhamar kota. India. Miner. Metall. Process.
2002, 19, 145–160. [CrossRef]
41. Aslan, N.; Fidan, R. Optimization of Pb flotation using statistical technique and quadratic programming.
Sep. Purif. Technol. 2008, 62, 160–165. [CrossRef]
42. Mehrabani, J.V.; Noaparast, M.; Mousavi, S.M.; Dehghan, R.; Ghorbani, A. Process optimization and modelling
of sphalerite flotation from a low-grade Zn-Pb ore using response surface methodology. Sep. Purif. Technol.
2010, 72, 242–249. [CrossRef]
43. Kwak, J.S. Application of Taguchi and response surface methodologies for geometric error in surface grinding
process. Int. J. Mach. Tools. Manuf. 2005, 45, 327–334. [CrossRef]
44. Awe, S.A.; Khoshkhoo, M.; Kruger, P.; Sandstrom, A. Modelling and process optimisation of antimony
removal from a complex copper concentrate. Trans. Nonferrous. Met. Soc. China. 2012, 22, 675–685. [CrossRef]
45. Bezera, M.A.; Santelli, R.E.; Oliveira, E.P.; Villar, L.S.; Escaleira, L.A. Review response surface methodology
(RSM) as a tool for optimization in analytical chemistry. Talanta 2008, 76, 965–977. [CrossRef]
46. Rakić, T.; Kasagić-Vujanović, I.; Jovanović, M.; Jančić-Stojanović, B.; Ivanović, D. Comparison of full factorial
design, central composite design, and Box-Behnken design in chromatographic method development for the
determination of fluconazole and its impurities. Anal. Lett. 2014, 47, 1334–1347. [CrossRef]
47. Box, G.E.P.; Hunter, J.S. Multi-factor experimental design for exploring response surfaces. Ann. Math. Stat.
1957, 28, 195–241. [CrossRef]
48. Asadi Kafshgari, L.; Ghorbani, M.; Azizi, A.; Agarwal, S.; Gupta, V.K. Modeling and optimization of Direct
Red 16 adsorption from aqueous solutions using nanocomposite of MnFe2 O4 /MWCNTs: RSM-CCRD model.
J. Mol. Liq. 2017, 233, 370–377. [CrossRef]
49. Hassanzadeh, A.; Karakas, F. The kinetics modeling of chalcopyrite and pyrite, and the contribution of
particle size and sodium metabisulfite to the flotation of copper complex ores. Particul. Sci. Technol. 2017, 15,
455–461. [CrossRef]
50. Hassanzadeh, A. A survey on troubleshooting of closed-circuit grinding system. Can. Metall. Q. 2018, 57,
328–340. [CrossRef]
51. Hassanzadeh, A.; Hoang, D.H.; Brockmann, M. Assessment of flotation kinetics modeling using information
criteria; case studies of elevated-pyritic copper sulfide and high-grade carbonaceous sedimentary apatite
ores. J. Disper. Sci. Technol. 2020, 41, 1083–1094. [CrossRef]
52. Gholami, H.; Rezai, B.; Mehdilo, A.; Hassanzadeh, A.; Yarahmadi, M.R. Effect of microwave system location
on floatability of chalcopyrite and pyrite in a copper ore processing circuit. Physicochem. Probl. Miner. Process.
2020, 56, 432–448. [CrossRef]
Minerals 2020, 10, 660 19 of 19

53. Raykov, T.; Marcoulides, G.A. An Introduction to Applied Multivariate Statistics; Routledge: New York, NY,
USA, 2008.
54. Niedoba, T.; Pi˛eta, P. Applications of ANOVA in mineral processing. Min. Sci. 2016, 23, 43–54.
55. Niedoba, T.; Surowiak, A.; Pi˛eta, P. A multidimensional analysis and modelling of flotation process for
selected Polish lithological copper ore types. E3S Web Conf. 2017, 18, 1–6. [CrossRef]
56. Jamróz, D.; Niedoba, T.; Pi˛eta, P.; Surowiak, A. The use of neural networks in combination with evolutionary
algorithms to optimise the copper flotation enrichment process. Appl. Sci. Basel 2020, 10, 3119. [CrossRef]
57. Jorjani, E.; Barkhordari, H.R.; Tayebi Khorami, M.; Fazeli, A. Effects of aluminosilicate minerals on
copper-molybdenum flotation from Sarcheshmeh porphyry ores. Miner. Eng. 2011, 24, 754–759. [CrossRef]
58. Park, C.H.; Jeon, H.S. The effect of sodium silicate as pH modifier and depressantin the froth flotation of
molybdenite ores. Mater. Trans. 2010, 51, 1367–1369. [CrossRef]
59. Fuerstenau, M.C.; Miller, J.; Kuhn, M.C. Chemistry of Flotation, Society of Mining Engineers of the American
Institute of Mining; Metallurgical and Petroleum Engineers, Inc.: New York, NY, USA, 1985.
60. Shahcheraghi, S.H.; Abdollahy, M.; Khalesi, M.R.; Hossein, S.M. A study on the effect of some operating
parameters on the flotation kinetics of copper sulfide minerals. In Proceedings of the XXVI International
Mineral Processing Congress (IMPC), New Delhi, India, 24–28 September 2012.
61. Dhar, P.; Thornhill, M.; Rao Kota, H. Investigation of copper recovery from a new copper ore deposit (Nussir)
in northern Norway: Dithiophosphates and xanthate-dithiophosphate blend as collectors. Minerals 2019,
9, 146. [CrossRef]
62. Roy, S.; Datta, A.; Rehani, S. Flotation of copper sulphide from copper smelter slag using multiple collectors
and their mixtures. Int. J. Miner. Process. 2015, 143, 43–49. [CrossRef]
63. Fornasiero, D.; Filippov, L.O. Innovations in the flotation of fine and coarse particles. J. Phys. Conf. Ser. 2017,
879, 012002. [CrossRef]
64. Hassanzadeh, A.; Firouzi, M.; Albijanic, B.; Celik, M.S. A view on determination of particle–bubble encounter
using analytical, experimental and numerical methods. Miner. Eng. 2018, 122, 296–311. [CrossRef]
65. Somasundaran, P.; Wang, D.Z. Solution Chemistry: Minerals and Reagents; Elsevier Press: Amsterdam,
The Netherlands, 2006.
66. Liu, C.; Feng, Q.; Zhang, G. Effect of ammonium sulfate on the sulfidation flotation of malachite.
Arch. Min. Sci. 2018, 63, 139–148.
67. Gu, S.; Fu, B.; Ahn, J.W. Simultaneous removal of residual sulfate and heavy metals from spent electrolyte of
lead-acid battery after precipitation and carbonation. Sustainability 2020, 12, 1263. [CrossRef]
68. Fairthorne, G.; Fornasiero, D.; Ralston, J. Effect of oxidation on the collectorless flotation of chalcopyrite.
Int. J. Miner. Process. 1997, 49, 31–48. [CrossRef]
69. Grano, S.R.; Sollaart, M.; Skinner, W.; Prestidge, C.A.; Ralston, J. Surface modifications in the
chalcopyrite–sulphite ion system, I. Collectorless flotation, XPS and dissolution study. Int. J. Miner. Process.
1997, 50, 1–26. [CrossRef]
70. Grano, S.R.; Cnossen, H.; Skinner, W.; Prestidge, C.A.; Ralston, J. Surface modifications in the
chalcopyrite-sulphite ion system, II. Dithiophosphate collector adsorption study. Int. J. Miner. Process. 1997,
50, 27–45. [CrossRef]

© 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/).

You might also like