You are on page 1of 31

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/337396401

Flotation of Iron Ores: A Review

Article  in  Mineral Processing and Extractive Metallurgy Review · November 2019


DOI: 10.1080/08827508.2019.1689494

CITATIONS READS

64 5,326

6 authors, including:

Natalia Parra Alvarez V. Claremboux


Michigan Technological University Michigan Technological University
4 PUBLICATIONS   65 CITATIONS    9 PUBLICATIONS   167 CITATIONS   

SEE PROFILE SEE PROFILE

Surendra Kawatra
Michigan Technological University
277 PUBLICATIONS   3,416 CITATIONS   

SEE PROFILE

All content following this page was uploaded by Surendra Kawatra on 19 October 2020.

The user has requested enhancement of the downloaded file.


Mineral Processing and Extractive Metallurgy Review
An International Journal

ISSN: 0882-7508 (Print) 1547-7401 (Online) Journal homepage: https://www.tandfonline.com/loi/gmpr20

Flotation of Iron Ores: A Review

Xiaolong Zhang, Xiaotian Gu, Yuexin Han, N. Parra-Álvarez, V. Claremboux &


S. K. Kawatra

To cite this article: Xiaolong Zhang, Xiaotian Gu, Yuexin Han, N. Parra-Álvarez, V. Claremboux
& S. K. Kawatra (2019): Flotation of Iron Ores: A Review, Mineral Processing and Extractive
Metallurgy Review, DOI: 10.1080/08827508.2019.1689494

To link to this article: https://doi.org/10.1080/08827508.2019.1689494

Published online: 19 Nov 2019.

Submit your article to this journal

Article views: 253

View related articles

View Crossmark data

Citing articles: 2 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=gmpr20
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW
https://doi.org/10.1080/08827508.2019.1689494

Flotation of Iron Ores: A Review


Xiaolong Zhanga, Xiaotian Gua, Yuexin Hana, N. Parra-Álvarezb, V. Clarembouxb, and S. K. Kawatrab
a
School of Resources and Civil Engineering, Northeastern University, Shenyang, PR China; bDepartment of Chemical Engineering, Michigan
Technological University, Houghton, Michigan, USA

ABSTRACT KEYWORDS
A tremendous amount of research has been done on refining the flotation process for iron ore and Iron ore; flotation; desliming;
designing the reagents which go into it. This paper reviews the industrial practices and fundamental reagents; surface chemistry
research surrounding iron ore flotation. The advantages and disadvantages of direct flotation,
cationic reverse flotation, and anionic reverse flotation are reviewed. A novel stepped flotation
technique is discussed for the treatment of carbonate-rich iron ores. The necessity of desliming to
enable effective flotation is discussed in detail. Selective flocculation desliming is discussed in
particular detail as it is the primary deslime methodology used in industry. This paper also describes
a large variety of reagents commonly used or in development in each flotation route, covering
collectors, activators, depressants, dispersants, flocculants, and frothers. The impact of the zeta
potential and surface chemistry on flotation and deslime is also discussed. The aim of this paper
is to provide a well-detailed, well-referenced source for the current status of iron ore flotation, and
thus provide a useful guide to its future development and to further improve flotation performance
for iron ore.

1. Introduction for many applications (Dobby 2002). The key feature of column
According to the U.S. Geological Survey there are approximately flotation that was adapted by many more recent flotation cells is
170 billion tons of iron ore available in the world, representing the use of wash water to reduce entrainment (Dobby 2002). By
83 billion tons of metallic iron which could be extracted. The adding wash water to the froth layer, poorly hydrophobic mate-
majority of this iron ore is distributed between Australia (29.4%), rial can be rejected from the column overflow, increasing overall
Russia (14.7%), Brazil (13.5%), and China (12.4%) (USGS 2017). selectivity. Care must be taken when scaling up flotation col-
Among these four reserves, Russia and Brazil have higher iron umns between pilot and plant-scale, as the silica rejection must
grades, while China has a lower iron grade (Tang et al. 2019; Yu be maintained at a very high level but the froth stability is
et al. 2019, 2019b; Zhang et al. 2019a). lowered by the increased flowrate of froth wash water (Dobby
Flotation is widely used as the primary beneficiation techni- 2002). One major downside of flotation columns is that they are
que for non-magnetic iron ores, and is often used for further difficult to control compared to mechanical cells, as the visible
upgrading magnetic iron ores as well (Araujo, Viana and Peres surface of the column shows comparatively little information
2005; Ma 2012; Nakhaei and Irannajad 2018). Flotation is fun- about its performance. Another disadvantage of column flota-
damentally a surface selective separation that separates hydro- tion is axial mixing – a set of 4 columns were installed in an
phobic materials from hydrophilic materials. By using various operating plant running cationic reverse flotation, but due to
reagents, the hydrophobicity of mineral surfaces can be selec- axial mixing they did not perform well compared to the existing
tively altered, allowing for a wide variety of separations to be flotation cells and were eventually removed.
achieved (Carlson and Kawatra 2013; Iwasaki 1999; Rao 2013). In the beneficiation process flotation generally takes
Because the success of flotation is mainly dependent on the place after comminution and any other separations, includ-
ability of air bubbles to carry the hydrophobized material to the ing desliming, but before agglomeration and induration.
froth, it has significant advantages over gravity separation tech- Because flotation operates on wet slurries while agglomera-
niques with very fine (<150 μm) particles. Magnetic separation tion uses primarily solid material, a dewatering step is
can also be made to work at fine particle sizes, but using mag- necessary between the two. Different flotation routes can
netic separation alone will reject the non-magnetic portion of the also have an impact on the efficiency of the dewatering,
iron ore. While flotation is often more expensive than magnetic especially if performed via filtration, and on the subsequent
separation, flotation also recovers the non-magnetic iron ore agglomeration steps.
fraction, and this increased recovery is often enough to offset Over the years, a tremendous amount of work has been
the additional cost. done to refine the flotation process, along with the reagents
Flotation can be performed in mechanical cells or flotation and equipment used (Clemmer 1947; Filippov, Severov and
columns. Flotation columns have become common in industry Filippova 2014; Zhu 1994).

CONTACT V. Claremboux vjclarem@mtu.edu Department of Chemical Engineering, Michigan Technological University, Houghton, MI, USA
© 2019 Taylor & Francis Group, LLC
2 X. ZHANG ET AL.

2. Flotation routes for iron ore (Houot 1983; Iwasaki 1983). This flotation route is quite simple
and can be made effective without desliming. At high pH values
The flotation of iron ore was first investigated in America
however, the surfaces of the common gangue minerals and the
starting in 1931 (Iwasaki 1983, 1999; Uwadiale 1992). Over
iron ore behave quite similarly, limiting separation power.
time, several different flotation routes have been devised.
A characteristic example of direct flotation in alkaline media
These routes can be divided into 3 categories: direct flotation,
can be found in the Republic Mine. Republic Mine’s ore is
reverse anionic flotation, and reverse cationic flotation. Direct
a specular hematite with small amounts of magnetite and martite.
flotation refers to processes which float the iron ore into the
The gangue is a recrystallized chert containing sericite, grunerite,
froth product, whereas reverse flotation processes float the
cummingtonite, and chlorite. Figure 1 outlines the refining pro-
gangue materials into the froth to be discarded.
cess at Republic Mine. Key highlights of this flowsheet are:
Direct flotation and reverse anionic flotation were initially
developed by Hanna Mining and American Cyanamid in the
1930s and 1940s. These methods were implemented in (1) Two stages of desliming involving two cyclones, the
Minnesota and Michigan concentrators by the 1950s (Filippov, underflow of which is treated with tall oil in a series
Severov and Filippova 2014). Cationic reverse flotation was of conditioners.
simultaneously being developed by the U.S. Bureau of Mines, (2) Regrinding to liberate residual quartz and silicates
which eventually became the primary flotation route for the from the iron minerals.
United States and other countries in the west (Filippov, (3) Steam injection into the pulp conditioner, which is
Severov and Filippova 2014). Meanwhile, anionic reverse flota- used to progressively raise the temperature to boiling.
tion became popular in China.
This process can upgrade the specular hematite from an iron
grade of 36.5% to 65.4% while maintaining an 82.5% iron
recovery. The flotation collector consists of approximately
2.1. Direct flotation of iron ore
91% oleic and linoleic acids, 6% rosin acids, remainder unsa-
Several flotation media have been used for direct flotation ponifiable (Houot 1983).
processes with iron ore, including alkaline media, acidic In the 1960s, investigations began to understand direct
media, dispersing media, selectively flocculating media, and flotation in weakly acidic media. The advantage of an acidic
most combinations of the previous either in series or (where medium is that in acidic conditions the surface charges of iron
possible) simultaneously. Direct flotation is not well-suited for minerals and many common gangues can be more distinct
some gangue materials – in particular, sulfur, phosphorous, than in alkaline flotation. This yields a greater difference in
fluorine, and carbonate gangue materials behave too similarly flotability, allowing for better separations (Bunge, Morrow
to iron ore in direct flotation to achieve good separations (Rao and Trainor 1977). One example of the acidic medium
2004; Uwadiale 1992). approach would be Groveland Mine in the United States,
The first flotation route used in industry was direct flotation shown in Figure 2, which achieved a final iron grade of
in an alkaline medium. Soda ash was used to adjust the pH value 64.3% with a silica content of 6.5% (Houot 1983).
to between 9–10, and fatty acids were used as collectors. This Another direct flotation concentrator is the Qidashan concen-
process was used starting in the 1950s by the Humbolt concen- trator in China, put into production in the early 1980s (Yong
trator in America and the Donganshan concentrator in China 2005). This process is shown in Figure 3, built to concentrate

Figure 1. Overview of the Republic Mine beneficiation process using direct flotation (based on Houot 1983).
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW 3

Figure 2. Overview of the Groveland Mine iron ore beneficiation process, utilizing direct flotation in acidic conditions (based on Houot 1983).

Figure 3. Overview of the iron ore beneficiation process in the Qidashan concentrator (based on Yong 2005).

from an iron grade of 28.94% to 63.6% in a weakly acidic medium intensity magnetic separation, gravity separation, and cyclones.
(Fan, Jing and Zhu 2011; Zhang and Dai 2012). The flotation In China, a lot of work going into designing efficient desliming
collector was sodium petroleum sulfonate, with sulfuric acid methodologies started in the 1980s (Sun 2006). Perhaps the com-
used as a pH regulator. Interestingly, the Qidashan concentrator mon approach for desliming is selective flocculation and
had previously run alkaline medium direct flotation, but could dispersion.
only achieve grades of 59.5% (Fan, Jing and Zhu 2011; Zhang and Comparative studies of direct flotation in alkaline and weakly
Dai 2012). acidic media have been carried out by the Maanshan Mine
Unlike alkaline medium flotation, weak acidic media are highly Research Institute in China. In general, the final iron grade was
sensitive to the presence of slimes. This is because under acidic found to be 2–4 percentage points higher at the same recovery
conditions the opposing surface charges allow the ultra-fine par- using weakly acidic media. Additionally, filtration of the mildly
ticles of the slimes to coat the other materials, causing them to acidic concentrate is simple. However, due to the requirement
behave like the slimes in flotation. Thus, desliming is required for desliming and the high collector usage, direct flotation in
to achieve high degrees of separation with direct flotation in weakly acidic media ends up having a higher overall cost.
weakly acidic media. Due to the fine particle size of slimes most In summary, direct flotation, whether alkaline or acidic, is
traditional separation methods are ineffective, including high- attractive due to the simplicity of the cost and reagent systems. It
4 X. ZHANG ET AL.

is suitable for the separation of low-grade and weakly-magnetic and improve separation, but in the case of hematite ores this
iron ores with relatively coarse liberation sizes. However, direct also tends to require the addition of a depressant to prevent
flotation has difficulty reaching iron grades above 64%, and the the flotation of hematite.
simplicity of the reagent system is somewhat offset by the high Cationic reverse flotation was initially developed by the
quantity of reagent necessary compared to alternative flotation U.S. Bureau of Mines branch in Minnesota, and is by far the
processes. Direct flotation processes around the world have most widely used in the iron ore industry (Filippov, Severov
largely been replaced by reverse flotation processes. Direct flota- and Filippova 2014; Houot 1983; Iwasaki 1983). This process
tion in acidic conditions is capable of achieving more effective floats silica gangue minerals such as quartz under alkaline
separations but is also much more sensitive to slimes, while conditions (pH 9.5–10.5) using amine collectors and starch
direct flotation in alkaline conditions is limited by the similarity depressants (Peres and Correa 1996; Papini, Brandao and
of the surface charges on both the gangue and valuable minerals. Peres 2001; Filippov, Filippova and Severov 2010). This pro-
cess primarily sees use while processing hematite and goethite
ores – some magnetite ores are floatable without a depressant
2.2. Reverse flotation of iron ore
at lower pH values (around 8). The starch depressants prevent
Reverse flotation is presently the most popular methodology the collector from attaching to the iron oxides, allowing for
for concentrating iron ore. In reverse flotation the goal is to a high degree of separation. Cationic reverse flotation has
cause the gangue materials to move up and into the froth, been applied widely in the United States, Brazil, Canada,
which is precisely the opposite of direct flotation (Ma 2012; India, Russia, and Chile (Lima, Valadão and Peres 2013;
Filippov, Severov and Filippova 2014). Reverse flotation is Peres et al. 2009). Research in cationic reverse flotation for
broadly divided into two categories depending on the choice iron ore continues to be an active field.
of collector: cationic reverse flotation and anionic reverse A characteristic example of the reverse cationic flotation
flotation. route is shown in Figure 4, which outlines the process of
Tilden Mine in the United States. Tilden Mine utilizes selec-
2.2.1. Cationic reverse flotation tive flocculation and dispersion followed by flotation to sepa-
Cationic reverse flotation utilizes a cationic collector to float rate a low-grade hematite ore (Villar and Dawe 1975). The ore
and reject negatively charged silica. Silica surfaces hold nega- is initially ground to a passing size of 80% passing 25 μm and
tive charges at all but exceedingly acidic pH values, so theo- dispersed using sodium hydroxide along with a dedicated
retically cationic reverse flotation can be performed at a wide dispersant: originally sodium silicate, but more recently
range of pH values. In practice, alkaline pH values are often sodium polyphosphates and since 2015 sodium polyacrylates.
used to disperse valuable minerals and silica from each other Corn starch is used to flocculate the iron-bearing minerals in

Figure 4. Overview of the Tilden Mine iron ore beneficiation process, utilizing cationic reverse flotation (based on Houot 1983).
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW 5

the deslime thickener, and the overflow is removed as slimes. magnetic separation suffices to increase the iron grade of the
Deslime overflow at Tilden accounts for 15–30% of the total feed material to 35–40%, after which it undergoes another grind-
material loss and the loss of 5% of the iron-bearing minerals. ing step to 85% passing 44 μm to liberate the iron ore. This
The deslime underflow typically has a grade around 44%, but reground ore is then deslimed and floated via reverse cationic
can vary widely based on the water quality. The deslime flotation to produce a final concentrate with an iron grade of
underflow feeds the flotation stage, consisting of 1 rougher 66% (Oyarzun 2013).
and 4 scavengers. Each scavenger recycles its concentrate back In Russia, the Mikhailovsk plant (MMPP) processes an
to the initial flotation feed, while passing its froth product to unoxidized ferruginous quartzite. While most Russian iron
the next scavenger (DeVaney 1985). The fourth scavenger’s ore is processed from the Kursk Magnetic Anomaly (KMA)
froth product is considered the final tailing product. The and possesses significant magnetic properties making mag-
flotation collector is an unneutralized primary ether amine, netic separation ideal, the MMPP ore does not readily sepa-
and corn starch is again added as a depressant. The final rate under magnetic separation. The MMPP has developed an
product has an iron grade of 64–67%, contains 5% silica, integrated magnetic separation-flotation process for its ore,
while recovering approximately 75% of the iron. which utilizes wet magnetic separation followed by reverse
Another example of reverse cationic flotation is the Empire cationic flotation using an amino ester collector and
Mine in the United States. In Empire Mine, the flotation feed a causticized corn starch as an iron minerals depressant. The
is a magnetite concentrate prepared with 2-stages of low- final concentrate has an iron grade of approximately 69.7%
intensity magnetic separation with an iron grade of 63%. (Varicheva et al. 2017).
While additional stages could be used to increase the grade In India, most iron ores are only washed and classified
further, 3-stage low-intensity magnetic separation had been before sale. However, as alumina rejection and the processing
tried in Griffith Iron Mine in Canada to achieve 66.5% iron of slimes (in this case, particles smaller than 37 μm) or tailings
grade in magnetite from a 29% iron grade iron chert. have become more important (Jain et al. 2013), focus has
However, the final product still contained more than 6.5% returned to cationic reverse flotation, often in tandem with
silica. Thus, instead of adding more stages to magnetic separa- hydrocyclones or magnetic separation (Kumar et al. 2010).
tion, Empire Mine chose to use reverse cationic flotation to In China, research on cationic reverse flotation began in the
remove the problematic silica directly. This choice allowed for 1970s. In the 1980s, it was successfully implemented at the
a final concentrate iron grade of 69% with less than 3.5% Donganshan Sintering Plant, resulting in a 4 percentage point
silica. The flotation step had a 96% recovery, using an amine increase in iron grade – from 61.50% to 65.50% (Fan, Jing and
salt (ADMAG-83) collector. Similar results were found by Zhu 2011). However, there are fewer varieties of cationic and
Samarco Mine which processed a 53% iron grade hematite amine collectors available in China, and many of the available
ore, which was ground to 92% passing 0.15 mm and floated collectors have considerable drawbacks: too much froth, too
using mechanical flotation cells and column flotation. sticky froth, poor liquidity of the froth, poor selectivity overall,
Samarco Mine’s product is about 15.5 million tons per year and a high sensitivity to slimes. As such, the reverse cationic
of crude concentrate at an iron grade of 67.17% and contain- flotation route was not widely popular at the time. In more
ing only 1.1% silica. recent years, China has been vigorously promoting saving
In Canada, iron mining operations started in 1954 in costs, limiting energy consumption, and environmentally-
Schefferville, Quebec, Canada. At that time, no particular friendly practices to intensify its economy, and as such cationic
ore beneficiation was required, but by the end of the 1950s collectors have undergone rapid development. “Study on new
research began on upgrading an iron deposit with an iron technology and new reagent for iron-increasing and silica-
grade of 51% and a silica content of 15%. In 1973, the first reduction by cationic reverse flotation” has been undertaken
pellets were produced after beneficiation with reverse cationic and completed by the Anshan Iron and Steel Company in
flotation. In 1981 the Sept-Iles concentrator flotation circuit accordance with the National “Tenth Five-Year Plan” in China
consisted of 8 rougher cells, 3 cleaner cells, 4 first stage as one of the Science-Technology Projects (Song and Li 2008).
scavenger cells, and 4 second stage scavenger cells. It operated It is interesting that cationic reverse flotation has not been
at a pH of 10.5 and produced a final concentrate with an iron so successful in China. Instead, research there has focused
grade of 64% and containing 4.5% silica at a 93.5% iron more heavily on anionic reverse flotation. There is little con-
recovery (Gagnon 1981). crete data to explain this discrepancy, but perhaps it is
In Brazil, 10 different concentrators in the Iron Quadrangle explained by the availability of reagents or a difference in
utilize reverse cationic flotation to upgrade a banded hematite mineralogy. For the United States in particular, many iron
quartz ore. Some plants use only flotation to separate the ore, ore deposits have extremely fine liberation sizes (below
while other plants use a combination of high intensity magnetic 38 μm), which necessitates desliming to avoid issues with
separation followed by flotation. Slimes are usually treated with entrainment. Many desliming routes lead very naturally into
separate flotation circuits to be reclassified (Peres et al. 2009). cationic reverse flotation, as cationic reverse flotation typically
In Chile, Candelaria’s mine is a copper beneficiation plant benefits from the dispersing conditions established during
that produces tailings with a magnetic iron grade of 2–10%. deslime. Thus cationic reverse flotation is typically preferred
These tailings are processed by a nearby concentrator, Planta in these cases (Colombo 1980, 1986; Bruckard, Smith and
Magnetite, which mixes them with a magnetite ore from Mina Heyes 2015; Haselhuhn and Kawatra 2015a, 2015b).
de Los Colorados (a part of the CAO Valle del Huasco deposit, To summarize, cationic reverse flotation has several advan-
iron grade around 40%). After comminution and classification, tages when compared to direct flotation (Filippov, Filippova
6 X. ZHANG ET AL.

and Severov 2010; Ma, Marques and Gontijo 2011; Song and flotability, leading to a high degree of selectivity and separa-
Li 2008; Zhang and Dai 2012): tion. The anionic reverse flotation process has also shown
a good adaptability to mineral variations, especially for iron-
(1) Higher flotation rates and low-operating tempera- bearing silicate minerals.
tures, simplifying practical implementation. While 55% of the iron ore in China is referred to as Anshan-type
the reagent system is necessarily more complex than iron ore. 40% of these Anshan-type ores are low-grade hema-
direct alkaline flotation, it is not unduly so. tite ores with uneven mineral dissemination sizes along with
(2) The operation is typically simple and reliable. There complex mineral compositions (Han et al. 2006; Sun 2006;
are only a few key reagents in the reverse cationic Fan, Jing and Zhu). As a result of these factors, these ores are
flotation process, and as such the flotation process is difficult to upgrade. Over the past 50 years, investigations
easy to operate and readily adaptable to varying oper- have been undertaken by research institutes, universities,
ating conditions. and industries to develop effective processes to handle these
(3) Reverse flotation and gravity or magnetic separation ores. The mineralogy of the Anshan ores has led to the
readily complement each other, taking advantage of application of anionic reverse flotation techniques in several
differing physical mechanisms to achieve better concentrators throughout China.
separations than are possible with only one technique One of the major processes resulting from these repeated
or the other. investigations is depicted in Figure 5, which utilizes multi-
stage grinding, gravity separation, magnetic separation, and
However, there are also some considerable drawbacks (Filippov, anionic reverse flotation to achieve economical processing
Filippova and Severov 2010; Ma, Marques and Gontijo 2011; costs while producing a useable product. This process has
Song and Li 2008; Zhang and Dai 2012): been implemented at multiple large-scale concentrators
within China, including the Qidashan concentrator,
(1) The amine collector can still adsorb to the mineral Gongchangling concentrator, Sijiaying Mine, Yuanjiacun,
surface by complex adsorption, meaning that selec- and other. During the flotation stage of this process the quartz
tivity is often lower than comparable anionic reverse is activated using calcium oxide, an industrial modified fatty
flotation processes. acid (RA-715) is used as a collector, corn starch is used to
(2) Compared to other collectors, many commonly used depress the iron oxide, and the pH is adjusted to 11.5 using
amine collectors have both collecting and frothing prop- sodium hydroxide. The results of these concentrators are
erties. This results in high viscosity and poor fluidity of presented in Table 1 below.
the froth during flotation. Etheramines in particular The largest concentrator in China is the Yuanjiacun con-
have considerable frothing properties, and it may be centrator, owned by Tai Steel, which processes 22 million
desirable to investigate non-frothing alternatives so tons/year of raw ore. It is supplied by a reserve of approxi-
that entrainment can be more completely controlled. mately 1.2 billion tons of Anshan-type hematite ore. The
This is discussed in more detail in section 4.3. liberation size of this ore is approximately 90% passing
(3) The cationic amine collectors are toxic, and present 45 μm, making gravity separation ineffective. As such, the
an environmental hazard if improperly disposed of. process outlined in Figure 5 is poorly suited for handling
this ore. An alternative process was developed by the
Changsha Mining and Metallurgy Research Institute consist-
2.2.2. Anionic reverse flotation ing of: grinding, low-intensity magnetic separation, high-
Anionic reverse flotation uses anionic collectors to float and intensity magnetic separation, followed by anionic reverse
reject positively charged silica. While silica surfaces only flotation. This flowsheet is depicted in Figure 6, and has
naturally hold a positive charge at highly acidic pH values, achieved stable operation at Yuanjiacun. The reverse flotation
calcium and magnesium cations can specifically adsorb to it. stage at Yuanjiacun utilizes a newly developed modified fatty
These cations contribute the positive charge to the silica sur- acid referred to as MH. Otherwise, it is similar to previous
face and the reagents which provide these calcium or magne- processes: calcium oxide is used to activate the quartz, iron is
sium ions are termed activators as a result. depressed using corn starch, and pH is adjusted to 11.5 with
Anionic reverse flotation was initially developed in China sodium hydroxide. The final concentrate has an iron grade of
(Yuan, Han and Yin 2007). In practice anionic reverse flota- 66.95% and a final iron recovery of 72.62% from a run-of-
tion floats silica gangues by activating the silica with lime and mine hematite ore with an iron grade of 31.72% (Chen, Ge
then floating them with fatty acids as a collector at elevated and Yu 2005; Mao, Huang and Zhao 2005).
pH (11–12) while using starch to depress the iron-bearing Another ultrafine Anshan-type hematite-based ore deposit
minerals. The primary advantages of anionic reverse flotation is present at Qidong Mine. This deposit is the largest iron ore
are a decreased sensitivity to slimes, and a lower collector cost deposit in the Hunan Province in China, with 360 million
as the fatty acids are a major waste product from the paper tons of reserve. The liberation size is 95% passing 37 μm. At
industry (Forsmo et al. 2008; Zhao, Yahui and Yongdan this liberation size again gravity separation is ineffective.
2012). In practice it has been found that the calcium ions Additionally, flotation by itself is also insufficient to achieve
provided by lime both activate the siliceous minerals while a quality separation. Instead, due to the considerable forma-
allowing for strong depression of the iron-bearing minerals in tion of ultrafine silica slimes which interfere with both catio-
highly alkaline conditions. This results in a large difference in nic and anionic reverse flotation methodologies, desliming is
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW 7

Figure 5. Overview of the beneficiation process used for low-grade Anshan-type iron ores in China via anionic reverse flotation.

Table 1. Summary of concentrator input feeds and final product compositions Jing and Zhu 2011; Sis and Chander 2003). These taconites have
for selected concentrators utilizing anionic reverse flotation.
an iron grade of approximately 39% and are ground to 60%
Raw Ore Iron Concentrate passing 44 μm. This flotation was performed without de-slime,
Iron Grade Iron Grade Silica Iron Recovery using calcium chloride to activate the silica, and depressing the
Concentrator (wt%) (wt%) (wt%) (wt%)
iron oxides with dextrin (Gum 9072). The anionic collector in
Qidashan 29.16 67.67 3.85 71.65
Sijiaying 29.14 66.50 4.50 75.00 this process was Acintol FA2, which contains 97.8% tall oil, while
Gongchangling 28.78 67.79 4.10 81.99 the pH was maintained at 11.5 using sodium hydroxide. The
Yuanjiacun 31.72 66.95 NA 72.62
Qidong 30.32 Approx. 63–65% NA Approx. 65–70% resulting concentrate had an iron grade of 60.3%, silica content
NA: Not Available.
of 6.0%, and an iron recovery of 90.5%. However, despite these
fairly successful results, the process did not see widespread
application due to the high cost of reagents and difficulties
arising during concentrate filtration (Filippov, Severov and
necessary. Thus, a process consisting of multi-stage grinding Filippova 2014; Ge, Yu and Zhu 2009).
followed by selective flocculation was developed, as depicted Anionic reverse flotation can also be used to float apatite
in Figure 7. An iron ore concentrator has been built following gangue materials, to lower the phosphorous content of the
this scheme with a capacity of 3 million tons/year, with an resulting concentrate. For an apatite-bearing magnetite ore,
iron grade target of 63–65% while maintaining an iron recov- successful separation was achieved using a fatty acid mixture
ery of 65–70% (Hu, Zheng and Xiao 2010; Liu 2003). This (Altrac 1563) of 95–98% tall oil and 2–5% maleic anhydride as
development represents a major breakthrough in the handling a collector, and sodium silicate as a dispersant for the fine
of ultrafine hematite ores via anionic reverse flotation. particles and a depressant for the magnetite (Forsmo et al. 2008).
Since the 1960s anionic reverse flotation of low-grade hema-
tite ores has also been carried out in the United States, Canada,
and the Soviet Union (Filippov, Severov and Filippova 2014;
2.3. Stepped flotation
Sandvik and Larsen 2014). In the United States, the Hanna
Mining and U.S. Bureau of Mines laboratories jointly developed Stepped flotation processes have recently been proposed for
the anionic reverse flotation route to separate the taconites of the treating difficult carbonate-bearing iron ores (Shao 2013).
Lake Superior deposits (Bunge, Morrow and Trainor 1977; Fan, These are processes which combine multiple flotation routes
8 X. ZHANG ET AL.

Figure 6. Overview of the beneficiation process used by the Yuanjiacun iron ore concentrator, utilizing anionic reverse flotation (based on Chen, Ge and Yu 2005;
Mao, Huang and Zhao 2005).

into a single complementary methodology. Carbonate miner- barrier to utilization. For the siderite-bearing iron ores in
als, which in iron ore primarily refers to siderite, are present China, the process begins by floating the siderite in neutral
in up to 5 billion tons of China’s iron ore reserves (Luo et al. direct flotation using an anionic collector and depressing
2013; Song, Yuan and Wei 2015). In practice it has been the hematite using starch (Zhang et al. 2007; Song, Yuan
found that these minerals have middling flotabilities and read- and Wei 2015). Afterward, the silica is separated by anionic
ily generate slimes during grinding. These slimes then end up reverse flotation in strongly alkaline conditions using lime
coating the surfaces of both the hematite and the silica phases, as an activator (Zhang et al. 2007; Song, Yuan and Wei
resulting in considerable difficulties achieving good separa- 2015). This process, depicted in Figure 8, was put into
tions with flotation. In particular, some of the silica would be practice at the Donganshan Concentrator in China yielding
depressed by the siderite slimes, which can itself be depressed a concentrate with an iron grade of 63.02% and an overall
partially by starch, while some of the hematite would fail to be iron recovery of 63.77%. This two-step process mitigates
depressed due to the presence of the same slimes (Yin, Han the impact of the siderite on the reverse flotation, allowing
and Xie 2010; Yin et al., 2011). As such, carbonate-bearing for the efficient utilization of these carbonate-bearing ores.
iron ores were considered too difficult to process, and were
stockpiled at the cost of increased land usage and the resulting
3. Desliming
environmental pollution risks.
Stepped flotation processes were developed by the Desliming is the separation of ultrafine particles, typically but not
Northeastern University in China to breakthrough this always silica, from coarser particles in a slurry during mineral
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW 9

Figure 7. The Qidong concentrator uses multiple deslime and regrinding stages to separate an ultra-fine iron ore with anionic reverse flotation (based on Hu, Zheng
and Xiao 2010; Liu 2003).

processing. Slimes interfere with many aspects of any beneficiation The presence of fine slimes also complicates the filtration steps after
process resulting in higher reagent consumption and reduced flotation. Thus, a high content of fine slimes negatively impacts the
separation efficiencies. As such, desliming is necessary when work- flotation rate and selectivity, while increasing reagent usage (Lima,
ing with finely ground iron ores, especially before gravity separa- Peres and Marques 2012).
tion, magnetic separation, and flotation (Tripathy, Banerjee and Industrially, desliming is typically accomplished using either
Suresh 2015). a thickener or a cyclone, and is primarily targeted at removing
With regards to flotation, it has been found that flotation of iron particles primarily smaller than 20 μm (Thella, Mukherjee and
ores is most effective with particle sizes ranging from 10 μm to Srikakulapu 2012). The particle size cut for cyclones is defined by
100 μm. Within the context of reverse flotation particles smaller the cyclone geometry, pressure, and material being separated,
than 10 μm may be referred to as fine slime (Krishnan and Iwasaki with the 20 μm being typical for applications in iron ore. Deslime
1984). These particles may nonselectively and spontaneously aggre- thickeners can achieve finer cut sizes by carefully controlling the
grate in the slurry, or they may coat coarser particles making rise rate against the terminal velocities of the particles.
surface selective separations like flotation less efficient. Due to the
large surface areas presented by these ultrafine particles, they can
3.1. Conventional desliming
adsorb a disproportionately large quantity of reagents despite their
low mass. These fine particles can also adsorb to and monopolize The earliest report proclaiming the importance of desliming
the surface of the air bubbles, preventing them from attaching to was published in 1947 (Clemmer 1947) in the context of
other minerals (Edwards, Kipkie and Agar 1980; Iwasaki 2000). cationic reverse flotation of iron ores. The theory behind the
10 X. ZHANG ET AL.

Figure 8. Flowsheet for the beneficiation of a carbonate-containing iron ore using stepped flotation in Donganshan concentrator, using both direct and reverse
flotation methods (based on Zhang et al. 2007; SoSong, Yuan and Wei 2015).

behavior of slimes has been well-established as primarily an equilibrium concentration was maintained. In anionic reverse
electrostatic phenomenon stemming from heterocoagulation flotation focusing on floating the goethite using 0.1 mM sodium
(Fuerstenau, Gaudin and Miaw 1958; Iwasaki et al. 1962a; dodecylsulfate at pH 3, the presence of goethite slimes had little
Usui 1972). When slimes have different surface charge sign effect on the goethite recovery up to slime concentrations of
than the bulk material they are attracted to and coat the bulk 1 g/L. However, quartz, kaolinite, and bentonite clays had an
material, causing the bulk material to display the surface increasingly detrimental effect on goethite recovery, with even
character of the slime material during flotation. a 0.05 g/L concentration of bentonite slimes reducing goethite
Iwasaki (2000) illustrated the effects of slimes on cationic and recovery below 10%. At 1 g/L, quartz is associated with an
anionic reverse flotation. Goethite slimes were found to quickly approximately 55% goethite recovery and kaolinite with an
inhibit the flotation of quartz by an equilibrium concentration of approximately 15% goethite recovery. In both cationic and
0.1 mM dodecylamine at pH 6, with concentrations as low as anionic flotation these drops in recovery are associated with
0.05 g/L of slimes lowering quartz recovery below 15%. Kaolinite the slimes and the coarser particles having opposing charges,
and bentonite had no effect on quartz recovery in cationic and thus the slimes will coat the coarse particles. This is part of
flotation up to slime concentrations of 1 g/L so long as the the reason why anionic flotation typically takes place at such
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW 11

alkaline pH values: under such conditions both the silica gangue content of 4.75% requires a 4-minute residence time to achieve
and goethite surfaces are negatively charged, limiting the without desliming, but only a 3-minute residence time with
amount of slime coating which can occur (Chen and Peng desliming. Although in both cases the final concentrate iron
2018; Iwasaki 2000). grade and silica content were the same (at 67.9% and 4.75%
It has also been demonstrated that ultrafine particles smal- respectively), the iron recovery increased from 81% to 84% due
ler than 10 μm have a negative impact on cationic reverse to desliming. To achieve the same recovery without desliming
flotation in general (Lima 2001). This is largely attributed to requires increasing the reagent dosage and would result in
the considerable surface area of such fine particles causing a higher recovery of silica in the concentrate.
high reagent usage. Mohanty and Das (2010) investigated the modeling of
One comparative example of the effects of desliming can hydrocyclone-based slime upgrading using sodium hexameta-
be found at the Mikhailovsky plant in Russia. Removing the phosphate as a dispersant for iron slimes at Indian iron ore
sub-10 μm particles from the wet magnetic separation tailings mines. These slimes have a high iron content (approx. 58%
resulted in a modest improvement in the selectivity for hema- iron grade) and are primarily contaminated with finer silica
tite of the cationic reverse flotation step. Removing the sub- and alumina, along with some organics and water. They
20 μm particles increased the selectivity dramatically, but also predicted ideal parameters for the diameter of the hydrocy-
increased the hematite loss during desliming from 7.6% to clone’s vortex finder, the inlet pressure, the amount of dis-
16.6% (Filippov, Filippova and Severov 2010). persant, and the solids concentration for use with the iron ore
Ma, Marques and Gontijo (2011) reports an example of the slimes generated, with the goal of achieving 60% recovery and
effects of desliming on the particle size distribution, which 65% grade from the hydrocyclone. Though this is not strictly
reduced the sub-10 μm particle fraction from 7% to 2%, the speaking the same goal as desliming, the challenges presented
sub-20 μm particle fraction from approximately 12% to approxi- by the extremely fine particle sizes (90% passing 20 μm feed)
mately 6%, the sub-50 μm particle fraction from approximately remain significant.
45% to approximately 38%, and the sub-100 μm particle fraction Jena et al. (2015) performed an investigation on a similar
from approximately 80% to approximately 72%. Ma, Marques Indian iron ore slime using hydrocyclones to design
and Gontijo (2011) also demonstrates that anionic reverse flota- a flowsheet for the upgrading of these slimes to a usable
tion achieves better separation over the ultrafine fraction than iron concentrate. In this case, the slime was found to have
cationic reverse flotation, whereas cationic reverse flotation a 57.1% starting iron grade, and a magnetic separation plus
achieves better separations at coarser size ranges. At sub-10 μm hydrocyclone circuit was designed to upgrade it to 62.5% iron
particle size ranges, cationic reverse flotation recovers only 3–7% grade at 77.3% recovery. Again, quartz and alumina were the
of the iron in the concentrate, but anionic reverse flotation major gangue materials.
recovers between 20–60% of the same material. Both anionic Kruger, Naik and Naudé (2018) performed another inves-
and cationic reverse flotations reject the majority of the silica tigation along the same lines, taking an ultra-fine material
from the concentrate. The main advantage of anionic reverse from a SLon (WHIMS) concentrator and processing it into
flotation then is maintaining a high recovery of iron while low-silica iron product using largely desliming unit opera-
rejecting almost all of the silica even at very fine particle sizes. tions. In this case, approximately half of the silica needed to
As anionic reverse flotation is less sensitive to the presence be removed from a typical African iron ore to meet the target
of fine slimes, Houot (1983) has claimed that desliming is not grade, which they managed to achieve using a combination of
essential to the operation of anionic reverse flotation. hydrocyclones and flotation. Around 60% of the iron recovery
However, this claim has not yet been validated with experi- was immediately realized in the hydrocyclone circuit, and of
mental evidence. Ma (2012) has also reported that anionic the remaining 40% another half of that was recovered using
reverse flotation has been successful in China’s iron ore cationic reverse flotation.
industry in recent years, as the omission of the deslime step In summary, slimes have been shown to have a notable,
reduces costs but the lower sensitivity to fine slimes can allow negative impact on the quality and efficiency of flotation.
even a potential improvement in final product grade over Desliming, regardless of methodology, allows for the improve-
cationic reverse flotation. It should be recalled that Chinese ment of many iron ore separations if there are ultra-fine
concentrators frequently use multiple stages of magnetic particles present.
separation which may be responsible for the removal of
a significant quantity of the slimes. As previously mentioned
however, a large enough quantity of slimes is ultimately detri-
3.2. Selective dispersion and flocculation desliming
mental to both flotation routes, as all slimes will increase
reagent consumption. Selective dispersion and flocculation desliming is a particular
Wang (2009) has studied the impact of desliming operations deslime technique which relies on surface chemistry principles
at the Karara concentrator in Australia. The Karara concentrator to separate ultrafine particles. In particular, the components
processes a low-grade and high-silica magnetite ore with which are to be removed are strongly dispersed into the
a liberation size of 80% passing 35 μm via anionic reverse solution to prevent settling, while the components which are
flotation. Both pilot-scale and plant-scale investigations showed to be retained are selectively flocculated to increase settling. In
that desliming positively impacted the overall flotation effi- the context of iron ore, the dispersion is typically accom-
ciency. Wang (2009) showed that higher degrees of separation plished using sodium hydroxide or other strong bases, while
are achieved more quickly after desliming. In particular, a silica the flocculation is typically performed using starch.
12 X. ZHANG ET AL.

It has been found that one of the key factors leading to promise in controlling calcium concentrations in the pulp.
effective desliming is a high degree of dispersion among the Iwasaki (2000) also reports that precipitating the calcium
fine particles in the pulp (Huang et al. 2016). This state can be with carbonate or ultrasonic treatments can also be effective
achieved by operating at highly alkaline conditions, such as by for controlling the calcium concentration.
using sodium hydroxide, which results in all mineral surfaces Krishnan and Iwasaki (1984) have found that sodium
carrying negative surface charges. silicate is nonselective for dispersing goethite-quartz mixtures.
The extent of alkalinity required was investigated by Peres, At low silicate levels corn starch selectively flocculates
Lima and Araújo (2003) and Lima, Peres and Marques (2012) goethite, but at high silicate levels it is ineffective.
for nine samples collected at the CVRD (now Vale) mines of Weissenborn, Warren and Dunn (1994) carried out labora-
Alegria and Fabrica Nova in the Iron Quadrangle, Brazil. tory scale selective flocculation on sub-10 μm iron ore tailings
These results found that when the pH remained above 8, containing a primarily kaolinite gangue. They tested a variety
dispersion was largely unaffected. The amount of slimes by- of commercially available starches and polyacrylamides to
passing the desliming typically had a significant effect on upgrade a 46.6% iron grade feed material. Wheat starch was
flotation behavior, where a larger amount of slimes present found to be the most selective for this material, achieving
lowered the selectivity index for most samples. However, two a 57–58% iron grade and a 65–75% iron recovery.
samples are unaffected by the increased slimes concentration. It is worth emphasizing that the process water going into
For one sample, this may be attributed to an atypically high selective flocculation is a complex system of ions and reagents.
degree of dispersion compared to the other samples. Haselhuhn, Carlson and Kawatra (2012) investigated the
As a consequence of this behavior, it is suggested that effects of water chemistry on the selective flocculation and
dispersants such as sodium silicate or tripolyphosphate be dispersion of a hematite ore in an industrial concentrator
used during the cationic reverse flotation of iron ores. plant. The pH was shown to have the greatest impact on
However, conventional desliming would still result in exces- selectivity, which can be attributed to the pH value’s effect
sive iron loss for ores with very fine liberation sizes. Instead, on the surface charges of all the materials present. The pH
dispersion combined with selective flocculation is proposed to value in this facility is maintained above 10.5 during deslime
treat these ores (Colombo 1980; Colombo and Frommer and flotation to ensure that the silica and hematite remain
1976). By adding a dispersant and starch flocculant to the dispersed. The pH is then adjusted toward neutral before
desliming step, the iron losses can be mitigated. This style of filtration, which causes a significant change in the water
desliming has been implemented at Tilden Mine, Michigan, chemistry as ion solubilities change and the dolomite and
United States (Villar and Dawe 1975). limestone fluxes are added.
To achieve the best results, it is necessary that the ground Huang et al. (2016) reports that the selective flocculation
pulps be properly dispersed before the starch is added. The and dispersion process using a humate-based flocculant was
mineralogy and water chemistry of the slurry have applied to a Bayer process red mud in China. Bayer process
a significant impact on the optimal reagent dosages for selec- red mud typically has a relatively high iron grade, as its
tive desliming. Heerema, Lipp and Iwasaki (1982) investigated characteristic red color primarily stems from the presence of
this sensitivity by testing flotation with a 1:1 goethite-quartz fine hematite. The humate-based flocculant was shown to
mixture. An iron grade of 28.5% was found to be the lower exhibit good selectivity under alkaline conditions: iron recov-
limit of separation available, due to the natural desliming ery was 86.25 ± 1.31% and the concentrate iron grade was
properties of the dispersed material. In the absence of calcium 61.12 ± 0.10%. This was achieved at 2% solids, a flocculant
ions (Ca2+) corn starch results in very good selectivity. dosage of 30 mg/L, a sodium silicate dosage of 200 mg/L,
However, even the addition of a fairly small quantity a slurry pH of 10.0, and continuous mixing at 1000 rpm. It is
(0.001 M) of calcium ions results in greatly lowered selectivity: found that the humates selectively bridge the iron minerals,
the iron grades were reduced from up to 48.5% to at best 32%. resulting in large iron flocs which can settle comparatively
The conclusion is that controlling the calcium concentration rapidly.
via complexation or precipitation can significantly improve Multi-stage flocculation-desliming has also been imple-
deslime selectivity. mented at a small plant scale. The plant in question has
This can be accomplished with chemicals such as ethyle- a capacity of 150,000 tons/year and produces an iron concen-
nediaminetetraacetic acid (EDTA), sodium tripolyphosphate trate with an iron grade of 64.77% at a recovery of 78.65%
(STPP), or sodium hexametaphosphate (SHMP) (Arol and (Liu and Shi 2009).
Iwasaki 1987; Manukonda and Iwasaki 1987). Heerema,
Lipp and Iwasaki (1982) found that fully complexing the
4. Flotation reagents for iron ore
calcium with EDTA restores the selectivity of the separation,
but STPP is less effective and SHMP is actively detrimental to Flotation reagents are chemicals which can either change the
the separation. It appears that SHMP indiscriminately dis- surface properties of minerals, making them more or less
perses the pulp, resulting in poor separations overall. As susceptible to flotation, or the properties of the liquid such as
such it can also be concluded that desliming remains sensitive stabilizing the froth layer. The reagents form the critical back-
to some of these calcium complexes or other calcium com- bone of providing flotation circuits the control and efficiency
pounds which may be formed in the slurry. Iwasaki (2000) required to separate various minerals from each other. The
reported that montmorillonite (the principle component of most common reagent schemes typically involve a variety of
bentonite), which has a high cation exchange capacity, shows reagent roles including: collector, frother, depressant, activator,
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW 13

dispersant, and flocculant (Nakhaei and Irannajad 2018). In present a negative surface charge. Cationic collectors are
many cases one chemical may serve in multiple roles. For used primarily in cationic reverse flotation, which as pre-
example, starch is used as a depressant in flotation and viously mentioned is popular in the United States, Australia,
a flocculant in desliming. Additionally, many amine collectors Canada, Brazil, and others. Most cationic collectors are
have significant frothing properties, often negating the need for amines, and several types of amines have been used very
a dedicated frother. Due to the critical importance of these successfully in iron ore flotation (Araujo, Viana and Peres
reagents to the flotation process, a significant amount of prac- 2005; Liu et al. 2019a, 2019b).
tical work and theoretical research has gone into understanding The earliest amine collectors were fatty amines such as
their roles in flotation. dodecylamine, pioneered by their utilization in the U.S.
Bureau of Mines cationic flotation process. However, fatty
amines were ultimately supplanted due to their poor selectiv-
4.1. Collectors
ity, cohesive bubble formation, and low collection power at
The role of the collector is to adsorb to the mineral surface low temperatures. They were largely replaced with more solu-
while increasing its hydrophobicity. The hydrophobized sur- ble amines, particularly ether amines (Ma 2012).
face can then attach to air bubbles, carrying the collected Papini, Brandao and Peres (2001) performed a large number
mineral into the froth phase. Because most minerals are sig- of bench-scale flotation tests using ore from the Iron
nificantly hydrophilic, many collectors are designed to have Quadrangle, Brazil based on the most common amine collectors.
very hydrophilic functional groups on one end of their mole- In particular, the collectors tested were: Clariant’s Flotigam SA-
cular structure, while presenting a hydrophobic tail end on B, Flotigam T2A-B, Collector 075/94, HOE F2835-B, Flotigam
the other. Common hydrophilic functional groups include EDA-B, and Flotigam EDA-3B; Sherex’s MG-70-A5, MG-83-A,
carboxylic acids (-COOH), amines (-NH2), sulfonic acids (- and MG-98-A3; Pietschemicals’ ECAN 04D, Nb 104, and Nb
SO3H), hydroximic acids (-COH•NOH), and others (Fan, Jing 112; Quimikao’s Colmin C12; and Akzo’s Poliad A-3. The final
and Zhu 2011). Depending on the charge of the functional result found that ether monoamines (Flotigam EDA-B, EDA-
groups after disassociation in water, the collectors can be 3B, MG-70-A5, MG-98-A3, Colmin C12, and Poliad A-3) typi-
classified as anionic (e.g. most acids), cationic (e.g. amines), cally performed very well, ether diamines (HOE F2835-B
nonionic (no groups contribute to a charge when dissolved in and MG-83-A) were somewhat less effective, and fatty amines
water), or amphoteric (groups contribute a charge, but the (Flotigam SA-B, Flotigam T2A-B, and Collector 075/94) and
molecule retains a neutral net charge). Nonionic collectors condensates (ECAN 04D, Nb 104, and Nb 112) showed overall
may be mixed with cationic or anionic collectors, though poor selectivity for cationic reverse flotation of silica from iron
this is only infrequently used in iron ore flotations. ores. However, it is worth noting that the Iron Quadrangle
Amphoteric collectors can be used to boost the performance possesses a unique mineralogy – ether diamines are typically
of cationic or anionic collectors. more effective than monoamines on coarser ores (Nakhaei and
Additionally, it is in some cases worthwhile to co-dose Irannajad 2018).
both cationic and anionic collectors in the same process. More recently, Weng et al. (2013) proposed a novel quaternary
Though they cannot both be simultaneously active in solu- ammonium surfactant called M-302. M-302 contains ester bonds
tion, as the ionic interactions between the collectors will result and hydrocarbon tails. The synthesis route is provided in Figure 9,
in the complete utilization of one of them, co-dosing yields involving the reaction of adipic acid and N-(2,3-epoxypropyl)
a practical solution for the separation of certain refractory dodecyldimethylammonium chloride, the latter of which can be
flotation feeds. prepared by reacting a dodecyldimethylammonium salt with
epichlorohydrin. M-302 was initially used as a cationic collector
4.1.1. Cationic collectors for floating silicates from a Chinese magnetite ore, and it demon-
Cationic collectors dissociate in water to hold a net positive strated a stronger collecting power, wider effective temperature
charge, and are attracted to negatively charged surfaces. For range, and a higher rate of foam collapse than dodecylamine
this reason, they are primarily used to float minerals that chloride.

Figure 9. Synthesis of M-302 (Weng et al. 2013).


14 X. ZHANG ET AL.

Huang et al. (2014) introduced a gemini surfactant, ethane- recovered approximately 40% of the specularite at pH 12,
1,2-bis(dimethyldodecylammonium bromide) (EBAB) for use in suggesting that a depressant such as quebracho may be
the cationic reverse flotation of silica from a magnetic separation required to achieve optimal specularite separation. This result
concentrate supplied by Gongchangling Iron Mine of the was reviewed by Nakhaei and Irannajad (2018), who came to
Anshan Steel Company in China. EBAB was synthesized from a similar analysis and concluded that the CS-22 was an appro-
N,N,N’,N’-tetramethylethylenediamine with 1-bromododecane. priate collector for separations of magnetite and quartz.
The chemical structures of the surfactant and the conventional Chen et al. (2017) reports a novel cationic surfactant, tribu-
monomeric surfactant dodecylammonium chloride (DAC) are tyltetradecylphosphonium chloride (TTPC), which was found
shown in Figure 10. The bench-top flotation experiments using to have higher collection power and selectivity than dodecyl-
EBAB found that the concentrate iron grade could be increased triethylammonium chloride (DTAC) for the flotation of silica
to 70.58% with a silica content of 1.79% at an iron recovery of from magnetite. At equimolar dosages and at a pH of 8, TTPC
98.42%, while DAC could only achieve a concentrate iron grade recovered over 95% of the quartz and less than 5% of the
of 69.15% with a silica content of 2.65% at an iron recovery of magnetite, while DTAC recovered approximately 60% of the
95.78% (Huang et al. 2014). In summary, EBAB presented quartz and approximately 8% of the magnetite. Figure 11
a higher collecting power and superior selectivity when com- depicts the chemical structures of TTPC and DTAC. Both
pared to DAC. One notable property of EBAB is its high charge TTPC and DTAC effectively attach to the silica surface due
density stemming from the relative proximity of the two ammo- to the silica’s negative surface charge, but the active portion of
nium groups. Nakhaei and Irannajad (2018) also reviewed this TTPC (-P+(C4H9)3-) presents a stronger positive charge than
result, going into far more detail on the specific mechanism of the ammonium group (-N+(CH3)3-) in DTAC.
adsorption. Nakhaei and Irannajad (2018) also note that there is Sahoo et al. (2015) investigated the use of Aliquat-336,
very little detailed literature on the use of gemini surfactants in a quaternary ammonium salt used as a catalyst for phase
the flotation of iron ore. From the results with EBAB, it appears transfer and for the extraction of metals, as a cationic
to be a field worth pursuing further. collector. Aliquat-336 contains a mixture of C8 and C10
Wang and Ren (2005) compared three collectors for the carbon chains, primarily C8, along with nitrogen functional
flotation of quartz from magnetite and specularite: groups which allow it to act as a cationic collector. Sahoo
a combined quaternary ammonium salt (CS-22), dodecylamine et al.’s (2015) investigations focused on the separation of
chloride, and cetyltrimethylammonium bromide (CTAB). CS- a 38% iron grade banded hematite quartzite collected from
22 was prepared by mixing dodecyldimethylbenzylammonium the Hospet area, Kamataka, India. The final concentrate
chloride and dodecyltrimethylammonium chloride in a 2:1 was able to achieve 65% iron grade at a 60% iron recovery.
ratio. Wang and Ren (2005) showed that CS-22 had superior Sahoo et al. (2016) also investigated several ionic liquids
selectivity and collecting power compared to DDA or CTAB. based on imidazolium, ammonium, and pyridinium chemis-
CTAB was largely unselective at pH values above 6, though tries. Dodecylamine (DDA) and cetyltrimethylammonium
showed high selectivity at pH 2. DDA showed the best selec- bromide (CTAB) were used as collectors while soluble starch
tivity between pH 8 and 10, but recovered approximately half was used as a depressant in the ionic liquid solutions for
of the iron-bearing minerals in the froth. CS-22 recovered separating the banded hematite quartzite ores. Sahoo et al.
almost all (>90%) of the quartz when the pH was above 6, (2016) found that the ammonium and pyridium based collec-
while recovering less than 8% of the magnetite. CS-22 tors showed higher iron grade and recovery than the

Figure 10. Chemical structures of ethane-1,2-bis(dimethyldodecylammonium bromide) (EBAB) (a) and dodecylammonium chloride (DAC) (b) (Huang et al. 2014).

Figure 11. Chemical structures of (a) tributyltetradecyl-phosphonium chloride (TTPC) and (b) dodecyl triethyl ammonium chloride (DTAC) (Chen et al. 2017).
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW 15

imadizolium collectors, which is attributed to the presence of acid salts. Fatty acids are obtained from the saponification of
longer alkyl chains allowing for greater hydrophobicity. various vegetable or animal oils, which can be processed into
Nakhaei and Irannajad (2018) report that some common collectors suitable for iron ore flotation. The degree to which
collectors in commercial usage in Western plants include: the fatty acids are unsaturated plays an important role in iron
Aerosurf MG 98, Aerosurf MG584, Aerosurf MG70A3, ore flotation. As a general trend, fatty acids containing longer
Armac CES342, MG98A3, MG83A, Specialty FR, and FA2. carbon chains tend to be more powerful collectors than those
These compounds are largely ether amines used to float silica, containing shorter carbon chains, as they provide more
aside from Specialty FR and FA2 which are fatty acids which hydrophobicity to the surfaces they attach to. It had been
were used to float iron in Republic mine. thought that fatty acids with a higher degree of unsaturation
There is continuing work to develop new cationic collec- tend to be overall stronger collectors, but beyond oleic acid
tors, although ether amines remain dominant in industry this trend does not continue (Uwadiale 1992).
despite comparatively low selectivity. These ether amines Changsha Mining and Metallurgy Research Institute in
often have significant frothing properties, which can make China recently developed the RA series of anionic collectors
independently controlling the froth properties difficult. This (Luo et al. 2015). The RA series collectors are prepared by
is discussed further in section 4.3. chlorinating and oxidizing tar oil, forming a mixture of
chlorinated and unchlorinated fatty acids and oleic acid.
4.1.2. Anionic collectors These collectors provide good performance, are simple to
Anionic collectors dissociate in water to hold a net negative produce, inexpensive, relatively nontoxic, and can be made
charge, which is subsequently attracted to positive surfaces. from comparatively ubiquitous raw materials. The first devel-
There are a large variety of anionic collectors available for use opment in the RA series was RA-315, which was applied to
with iron ore and can be broadly divided based on their the Qidashan concentrate of the Anshan Steel Company in
chemistries, the most common of which are listed below China to upgrade it to 65.33% iron grade at 80.72% iron
(Zhang and Dai 2012): recovery. RA-315 has also been applied to Donganshan and
Gongchangling iron ores, which are comparatively difficult to
(1) Carboxylic acids and their salts (-RCOOH, -RCOONa, process, as well as to Tilden ore from the United States. In
-RCOOK, etc.). Common examples are fatty acids, e.g. general, RA-315 was found to be more economically efficient
oleic acids and sodium oleates. than other anionic collectors for these ores. RA-515, RA-715,
(2) Alkyl sulfonates and their salts (-RSO3H, -RSO3Na, and RA-915 were developed afterward, of which the latter two
-RSO3K, etc.), such as sodium petroleum sulfonate. are currently in industrial use for anionic reverse flotation.
RA-715 and RA-915 have consistently shown good collect-
Sodium petroleum sulfonate is prepared by sulfonating and ibility and selectivity, and their chemical structures are out-
subsequently saponifying petroleum byproducts. Sulfonation lined in Figure 12. The chlorine present acts as an electron
can proceed by reacting a terminal methylene group with SO3. absorbing group, which is believed to improve the chemical
Saponification occurs by the addition of NaOH. adsorption of the collector to alkali metal ions. A chelate can
The earliest iron ore flotations were direct flotation using be formed with RA-915 by introducing a hydroxyl group,
anionic collectors, typically fatty acids, petroleum sulfonates, or which can in turn physically adsorb to alkali metal ions.
much later hydroxamates (Zhu et al. 2016). These, along with Sinosteel Maanshan Mining Research Institute in China
resin acids, soaps, alkyl sulfates, and alkyl sulfonates are often used has developed a low-temperature resistant anionic collector,
to float iron ores, especially hematite (Liu, Zhang and Liu 2007). namely MH-80, by modifying fatty acids (Zhou 2017). MH-80
However, the depression of gangue minerals remains a difficult was applied in the Jianshan Iron Mine of Taisteel, China to
problem in direct flotation, and in more recent years anionic upgrade a magnetite concentrate to over 69% iron grade.
reverse flotation has become preferred in China’s major iron Abaka-Wood, Addai-Mensah and Skinner (2017) reported
mines. successful flotation of monazite and hematite with sodium
Since the 1950s Chinese scientists and technicians have oleate (SO), sodium dodecylsulfate (SDS), and an alkyl hydro-
studied anionic collectors, starting with fatty acids and fatty xamic acid (AH) for pH values above the isoelectric point.

Figure 12. Chemical structures of RA-715 and RA-915 (Luo et al. 2015).
16 X. ZHANG ET AL.

The monazite was typically more susceptible to flotation than 4.1.3. Mixed collectors
the hematite. The silica recovery in the flotation concentrate Mixed collectors are typically mixtures of anionic and cationic
depended on the collector used, with SO collecting virtually collectors which can have increased selectivity when com-
none of the silica and the other two collecting only a small pared to single collectors (Holland and Rubingh 1992; Rao
quantity of silica. SO was found to reach the highest monazite et al. 2011). Appropriately synergistic mixtures can reduce
recovery, while SDS recovered the greatest amount of iron collector consumption, which in turn reduces production
overall. SDS required higher pH values to achieve lower cost and environmental pollution.
quartz recoveries, but above pH 10 monazite and hematite Rao, Antti and Forssberg (1990) and Rao and Forssberg
recoveries dropped as well. For SO, optimal recovery was (1993, 1997) reported that mixtures of cationic and anionic
achieved at a pH of 9, while AH achieved its optimal separa- collectors could achieve better selectivity for silicate flotations
tions at slightly alkaline pH values of 7–9. than either alone. This phenomenon was attributed to the co-
Luo et al. (2015) reports the synthesis of α-bromodecanoic adsorption of the anionic collector to the silicate surface by
acid (CH3(CH2)7CHBrCOOH, α-BDA) for use as an anionic attaching to the tail of the cationic collector, which leads to an
collector with calcium chloride as an activator. The synthesis increase in overall hydrophobicity. Cationic reverse flotation
was performed by treating decanoic acid with bromine in the can similarly be improved by the addition of nonionic surfac-
presence of a small quantity of phosphorous trichloride as tants, such as fatty alcohols.
a catalyst at 80–90°C for 7 hours. Single mineral flotation tests According to Leal Filho and Rodrigues (1992) the mixture
were performed using calcium chloride as an activator over of a primary ether monoamine with oxyethylated nonylphenol
a range of pH values at a temperature of 16°C. α-BDA has at an approximately 4:1 ratio improved silica recovery in
been shown to perform quite well at the relatively lower cationic reverse flotation. The addition of the nonylphenol
temperature, but is dependent on the presence of calcium also decreases the surface tension of the froth, reducing the
ions to achieve a high quartz rejection in reverse flotation. overall froth depth.
Without calcium the maximum silica rejection was only Vidyadhar et al. (2002, 2003) reports that nonionic surfac-
11.9%, but at a calcium concentration of 0.4 mmol/L the tants reduce electrostatic repulsion between dodecylamine
rejection increased to 99.5%. cations during adsorption. This reduction allows for a closer
Cao, Zhang and Cao (2013) investigated the effects of packing of particles in the froth layer, reducing the required
modifying sodium oleate with sulfuric acid. The modified dosage of dodecylamine.
sodium oleate (MSO) was compared to unmodified sodium Pereira et al. (2006) suggests diesel oil to ether amine at
oleate (SO), and MSO was found to be more tolerant of a 4:1 ratio while using Tergitol TMN-10 as an emulsifier at
fluctuations in process pH, temperature, and reagent dosages. a 19:1 ratio of oil:emulsifier. When tested on a Brazilian iron
Both reagents were tested at a 600 g/t dosage on a magnetite ore at a 49.1% iron grade and containing 28.5% silica, an iron
concentrate from Qiucheng Iron Mine, with MSO achieving concentrate containing 0.7% silica was obtained at a 74.4%
a 64.79% iron grade with a silica content of 3.18% at 93.60% total iron recovery.
recovery while unmodified SO achieves a 63.77% iron recov- With regards to magnetite concentrates, Filippov,
ery with a much higher silica content of 4.23% and lower iron Filippova and Severov (2010) reports that mixtures of ether
recovery of 92.77% (Cao, Zhang and Cao 2013). These trials amines with either primary monoamines or alcohols are
were laboratory closed-circuit trials. effective for cationic reverse flotation. The mixture of collec-
Quast (2006) investigated the behavior of basic satu- tors was found to form a hydrophobic layer on amphiboles
rated fatty acids as anionic collectors with carbon chain and magnetite-silicate aggregates even in the presence of
lengths of 6 to 18. It was found that for the shorter chain starch. This allows for a thorough separation of silicate con-
fatty acids the predominant surface interaction was taining particles, allowing for concentrates containing up to
between the undissociated acid group and the mineral, 70.3% iron grade and less than 1.0% silica from a feed con-
while at longer chain lengths the dissociated anionic acid taining approximately 3% silica.
groups became more important. At longer chain lengths, Vidyadhar, Kumari and Bhagat (2012) investigated the
the fatty acids also became too insoluble to be effective recovery of hematite using sodium oleate in the presence
flotation reagents. At tetradecanoic acid the highest flota- and absence of dodecylamine at natural pH (6–7). This
tion recovery was found to occur at the saturation point of involved conditioning the sample for 5 minutes in a prepared
the acid, implying that the overall flotation is being solution of the reagent, and then transferring it to the float
directly limited by the limited solubility of the acid. The cell. In the absence of dodecylamine, increasing the sodium
longer chain acids were generally effective at more alkaline oleate concentration improved hematite recovery. In the pre-
pH ranges than shorter chain acids, with for example sence of dodecylamine the same trend is seen until the
octanoic acid primarily being effective below pH 7 and sodium oleate concentration exceeds the dodecylamine con-
dodecanoic acid primarily being effective above pH 5 centration at which point a sharp decrease in recovery is
(except at pH 9, which showed an unusually low recov- observed. This can be attributed to interactions between the
ery). Hexanoic, tetradecanoic, hexadecanoic, and octade- two excess collectors forming a complex precipitate with
canoic acids were comparatively ineffective collectors randomly oriented surface sites, which provides little to no
when compared at 1.0 kg/t dosages, and were unable to benefit to flotation. At concentrations lower than the dode-
achieve flotation recoveries above 55%. cylamine however, the mixed collector has superior
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW 17

performance, attributed to the incorporation of the oleate from the formation of Ca(OH)+ from the calcium hydroxide that
between amine attachment sites which helps mitigate their forms from the reaction of calcium oxide and water. When the pH
electrostatic repulsion from each other (Vidyadhar, Kumari is below 11, calcium cations primarily exist as Ca2+, and the
and Bhagat 2012). activation of quartz is comparatively weak. At higher pH values,
Tian et al. (2017) investigated the separation of ilmenite Ca(OH)+ becomes preferred, leading to a strong activation of the
from titanaugite using the anionic collector sodium oleate quartz. This activation is key to the high selectivity of the anionic
(NaOL), the cationic collector dodecylamine acetate (DAA), reverse flotation process.
and a mixture of the two (NaOL-DAA). The mixture was Note that most of these activators can interact with clay
prepared by adding the two reagents together and mixing materials such as bentonite and kaolinite (Chen and Peng
for 5 minutes before utilizing it as a collector. The order in 2018; Ma 2010; Ma and Bruckard 2010; Ma, Bruckard and
which the reagents were added was not found to impact the Holmes 2009). With regard to flotation, these interactions are
flotation results. It was found that ilmenite and titanaugite typically undesirable: the metal cations added by CaO or NaOH
could be separated using the mixture within the pH range of tend to cause these clays to become more receptive to adsorbing
5–7. Reagent consumption using the mixture can be reduced to the hematite, making separations more difficult (Chen and
by up to half as compared to NaOL alone. This mixed collec- Peng 2018). Though the highly alkaline pH of cationic reverse
tor combines the superb flotation performance of DAA with flotation does much to mitigate this effect, if neutral or acidic
the relatively high selectivity of NaOL. salts are added it should be kept in mind. Similarly, the water
Yang et al. (2019) demonstrated that cetyltrimethylammo- chemistry can have a significant impact on the flotation of clay-
nium bromide (CTAB) mixed with sodium oleate (NaOL) also rich ores.
achieved better separations than either individually. The mixed
reagent was prepared by mixing the reagents separately before 4.2.2. Depressants
addition. Once added to the flotation cell the slurry and the Depressants either prevent the attachment of the collector or
reagents were mixed together for 5 minutes. In particular decrease the hydrophobic nature of the minerals, preventing
0.2 mM concentration of a mole ratio of 2:1 CTAB:NaOL at them from being transported to the froth. In reverse flotation
pH 5.5 achieved an iron grade of 57% at a recovery of 79%, while focusing on floating silica, the iron ore needs to be depressed.
neither collector individually at similar or even much higher Common iron oxide depressants are starch, guar, dextrin, and
dosages could exceed a 65% recovery. CTAB achieved 53% carboxymethylcellulose (CMC) (Nakhaei and Irannajad 2018).
iron grade at 42% recovery at a 0.12 mM dosage, while NaOL Starches can be extracted from several vegetable species: corn,
achieved a 46% grade at 60% recovery at a 3.0 mM dosage. cassava, potato, wheat, rice, arrowroot, and so on. These
starches are the primary depressants used for the depression
of iron ores (McDonald and Kawatra 2017; Nakhaei and
4.2. Regulators
Irannajad 2018; Turrer and Peres 2010). Corn starch is per-
4.2.1. Activators haps the most widely used of all of these, owing to its high
Activators interact with the mineral surface to promote the availability and good results in flotation. However, starch
interaction of the mineral with the collector. While in some choice has historically been primarily motivated by cost effi-
cases the mineral surface will naturally interact with the collector, ciency (Nakhaei and Irannajad 2018), rather than any parti-
several minerals require the addition of an activator for collec- cular quality metric. Most starches will adequately depress
tion to proceed (Rao 2013). This is quite common in reverse iron ore under some condition.
anionic flotation as, at the pH ranges involved, the anionic The primary selective binding mechanism of starch may be
collector and the mineral surfaces to be floated both tend to due to the very similar spacing between the hydroxyls of the
hold negative charges. Silicates tend to require the presence of amylose and amylopectin pyranose rings. In crystalline amy-
polyvalent metal cations, often Ca2+, Mg2+, or Fe2+ to activate lose, which is likely strongly indicative of the structures of
them. The pH ranges for which flotation can take place is both dissolved amylose and amylopectin, the spacing between
dependent on the presence of soluble metal hydroxide cations: the nearest adjacent hydroxyl groups is approximately 293 pm
too low of a pH and the bare metal cation does not appropriately and the closest spacing between the oxygen of the pyranose
adsorb to the surface, too high of a pH and the metal hydroxide ring and a hydroxyl group is 283 pm (Takahashi, Kumano
precipitates instead of adsorbing. Fuerstenau and Palmer (1976) and Nishikawa 2004). Meanwhile, in hematite, the spacing of
report these values in the presence of 0.1 mM sulfonate activator surface iron sites and correspondingly of surface hydroxyl
and 0.1 mM of metal salt: for iron (iii) the metal hydroxide sites is 285 pm (Haselhuhn 2015; Pradip 1994; Ravishankar,
cation begins forming at pH 2.9 and ceases to allow sulfonate Pradip and Khosla 1995). This similar chemical spacing would
adsorption at 3.8, for aluminum the metal hydroxide begins suggest that pyranose-like structures may be an important
forming at pH 3.8 and ceases to allow sulfonate adsorption at functional group to consider when designing new hematite
8.4, for lead the metal hydroxide begins forming at 6.5 and depressants.
ceases to allow sulfonate adsorption at 12.0, for manganese (ii) Typical industrial corn starches contain a starch fraction
the metal hydroxide begins forming at 10.9 and ceases to allow consisting of amylose and amylopectin, along with proteins,
sulfonate adsorption at 11.7, and for calcium the metal hydro- oils, fibers, mineral matter, and moisture. A typical industrial
xide forms and allows adsorption at all pH values above 12. corn starch contains 20–30% amylose, 70–80% amylopectin, and
Calcium oxide (CaO) is widely used to activate silica in iron ore less than 1% lipids and proteins. The active component of corn
flotation. The activation effect of calcium oxide is believed to stem starch when used as a depressant is the amylose and the
18 X. ZHANG ET AL.

amylopectin (Fuerstenau, Jameson and Yoon 2007). Iwasaki and and found that soluble starch was the most effective. Otherwise,
Lai (1965) observed that starches with higher amylopectin con- each starch was found to be similar. Potato starch achieved the
tent adsorb more readily to hematite surfaces. Amylose also least hematite depression, then rice starch, corn starch, and
adsorbs to hematite, but the effect of amylose on flotation and solubilized starch in turn showed increasing hematite depres-
flocculation is comparatively limited (Pinto, Araujo and Peres sion. Rice and potato starch depressed less quartz than corn or
1992). soluble starch, but the overall difference is at most 8 percentage
Despite practical evidence showing that most starches are points with a dodecylamine collector. The key highlights are that
effective for flotation, there was concern that proteins present maximum adsorption of each starch occurs in the pH range of
within the starch may interfere with the depression of iron 5–9, and that despite having the minimum absorption soluble
ores. However, Peres and Correa (1996) found that zein, the starch consistently depressed more hematite. While the depres-
most common protein found in corn starch, is an excellent sion of hematite increased with increasing starch adsorption, no
iron ore depressant on its own and outperforms pure amylose trend was found correlating starch adsorption with quartz
or amylopectin in lowering quartz recovery. However, gluten, depression.
which has a comparatively high protein content compared to Kar et al. (2013) also investigated the recovery of iron from
starch, is a significantly less effective depressant than corn low-grade banded hematite quartzites from Karnataka in
starch or zein. India using these same four starches as depressants. Using
Pavlovic and Brandao (2003) investigated the depressant dodecylamine at 32 g/t as the collector, the iron grade of the
effects of corn starch, amylose, amylopectin, glucose, and hematite concentrate product initially increases with increas-
maltose on hematite and quartz. Pavlovic and Brandao ing starch dosage, but eventually the depression of the quartz
(2003) found that amylose, amylopectin, and corn starch all limits further separation. Soluble starch was found to give the
effectively attach to the quartz surface, though amylose more most effective flotation, reaching an iron grade of 63.8% with
so. They further found that all of these compounds effectively an iron recovery of 89.6% with a starch dosage of 400 g/t.
depress hematite. However, amylose is the least effective at Ma (2008) investigated the effect of Na+ and K+ ions on the
depressing quartz, and amylopectin and corn starch were far adsorption of starch to quartz. In general, it was found that at
more effective for that. This is attributed to amylose not being low concentrations the actions of sodium and potassium did
a very effective flocculant as it is a linear sugar, whereas not differ significantly. However, at concentrations above
amylopectin is a more effective flocculant because its branch- 10 mM, an elevated concentration of potassium ions greatly
ing structure allows for more effective bridging. Thus, despite increased the adsorption of starch onto quartz at pH 10.5. It
the considerably lower absorption of amylopectin, it was far was determined that the solvation of starch onto quartz
more effective at depressing the silica. However, in reverse required energy, and only proceeded once a certain amount
flotation the poor quartz depression of amylose may be desir- of starch was added. The addition of potassium cations reduced
able, as amylose is still effective in depressing hematite allow- that energy by disrupting the structure of the hydrogen bonds
ing for the separation to proceed. Glucose and maltose, which in water, allowing starch to adsorb far more readily.
are also linear sugars, were found to have even less of an effect Lelis, Da Cruz and Lima (2019) investigated the effect of
on quartz flotation than amylose, suggesting that to depress Ca2+ and Cl− on the adsorption of starch to quartz and
quartz the flocculation behavior may be more important than hematite in an itabiritic iron ore. The addition of calcium
the surface modification. cations tended to increase the adsorption of starch to both
Ma and Bruckard (2010) investigated the depressant effect of minerals, resulting in increased depression in both cases. This
corn starch on kaolinite. They reported that at high pH values is typically undesirable, as increased quartz depression is
starch has almost no affinity for kaolinite in distilled water, synonymous with decreased quartz rejection. The addition
which means that kaolinite does not inherently interfere with of ethylenediaminetetraacetic acid (EDTA) to precipitate the
selective flocculation or depression of hematite or magnetite any calcium fully restored the flotability of the quartz. The pre-
more than any ultrafine slimes ordinarily would at these pH sence of chloride however resulted in the formation of
values. Below a pH of 9, however, the surface charge of kaolinite a lasting complex with the hematite, hindering starch adsorp-
becomes more positive and starch adsorption increases. The tion. Thus, after the addition of CaCl2, the adsorption of
affinity between starch and kaolinite is also dependent on ionic starch to hematite was partially compromised, even with the
strength (Ma and Bruckard 2010; Ma, Bruckard and Holmes addition of EDTA to capture the calcium cations.
2009). With a NaCl concentration of 1 mM at pH 10.5, the Zhang et al. (2017) reported a novel hematite depressant,
adsorption of starch on kaolinite doubled over the adsorption the macromolecule polymaleicanhydridetriethylenetetra-
in distilled water. The presence of other metal cations is also mine (PMTA). The synthesis route of PMTA is shown in
known to influence the structure and attachment properties of Figure 13. Using dodecylamine as the collector at pH 11,
kaolinite, modifying its affinity for hematite as well as starch PMTA achieved a flotation grade of 65.24% and a recovery
(Chen and Peng 2018). Thus, while it is theoretically possible to of 98.29% while starch achieved a flotation grade of 64.93%
separate kaolinite from hematite using starch as a depressant, and a recovery of 95.35%.
doing so on a plant-scale requires very careful control of the Aside from starch, carboxymethylcellulose (CMC) has
water chemistry. shown potential as an iron ore depressant. Testing modified
Kar et al. (2013) investigated the effects of different types of CMCs with iron ore found that high iron grades (>68%) could
starch on cationic reverse flotation using dodecylamine. Soluble be achieved in the concentrate (Poperechnikova et al. 2017).
starch, corn starch, rice starch, and potato starch were compared, However, CMCs typically lead to greater iron losses to the
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW 19

Figure 13. The synthesis route of polymaleicanhydridetriethylenetetramine (PMTA) (Zhang et al. 2017).

tailings as compared to starch alone. As CMC is also more There are several reagents that may be used in selective
expensive than starch, the loss of iron makes it difficult to flocculation and dispersion, depending on the mineralogy.
recommend universally. These reagents adsorb to the mineral surfaces and induce either
Turrer and Peres (2010) performed a comparative analysis of dispersion or flocculation in the ore particles (Haselhuhn and
several depressants, including 6 CMCs, 3 lignosulphonates, 1 Kawatra 2015a, 2015b). Figure 14 shows some example disper-
guar gum, and 4 humic acids. However, only CMC and guar sants to mitigate the effects of slimes during this process – of
gum were able to reach the same flotation performance as starch, these, sodium silicate, sodium hydroxide, and sodium carbonate
which is attributed to the presence of the glucopyranose ring. are the most commonly used in industry (Haselhuhn 2015).
However, blends of these polymers achieved promising results. Ma (2011) investigated the dispersion of kaolinite gangue
Humic acids (HA) have been found to preferentially adsorb materials in the context of the cationic reverse flotation of
to certain iron-bearing minerals (Illés and Tombácz 2003; Lai iron ore. Ma found that while sodium silicate is widely used
and Chen 2001; Ramos-Tejada et al. 2003; Tombacz et al. 2004). for dispersing kaolinite, it is only effective at acidic pH values.
This preferential adsorption can lead to a depressant effect on Since cationic reverse flotation often takes place at pH values
these minerals. Humic acids are high molecular weight polymers above 9.5, this is less than ideal. However, sodium polypho-
which are internally apolar but present several hydrophilic sphate and polyacrylic acid were able to disperse kaolinite at
groups (primarily carboxylic acids and phenolic groups) on pH values up to 10.5. The adsorption of polyacrylic acid to
their exterior (Chi and Amy 2004; Terashima, Fukushima and kaolinite can be improved further by the addition of CaCl2 or
Tanaka 2004). Dos Santos and Oliveira (2007) studied the use of MgCl2 at high pH values.
humic acid as an iron ore depressant for a 1:3 quartz:hematite For flocculants, the most commonly used are starch, poly-
mixture in flotation. The resulting concentrate was 86.0% hema- acrylamide, and humate. Macromolecule flocculants have seen
tite at >90% hematite recovery. greater interest in recent years, however. This is in part due to
the continued growth and development of the industry, and
4.2.3. Dispersants and flocculants in part due to the increasing demand for environmentally
Increasing demand for iron ore has necessitated investiga- sustainable processes.
tion into the processing of low-grade, very fine iron ores, Starch is a natural polymer consisting primarily of amylose
the extraction of iron from iron-bearing slimes, and the and amylopectin. Starches display a particular affinity toward
extraction of iron from existing tailings. Selective dispersion iron minerals, which plays a role in its use as a depressant as
and flocculation is the primary process for separating these discussed earlier, but also plays a key role in its action as
fine materials. Selective flocculation depends on the prefer- a flocculant. Starch’s affinity for iron ore stems from its ability
ential adsorption of organic flocculants onto specific to form complexes and hydrogen bonds with the iron ore
mineral particles. These flocculants cause the particles to surface, and as such it has been utilized in commercial flota-
settle quickly, while dispersants are used to keep the tion and selective flocculation processes (Panda et al. 2013;
remaining material largely suspended. To maximize the Weisseborn, Warren and Dunn 1995). The Tilden concentra-
effectiveness of this process, the pulp needs to be fully tor in the United States has been very successful utilizing corn
dispersed before any flocculant is added. This process has starch for selective flocculation of iron ore.
been quite successful for the extraction of valuable minerals Weissenborn (1996) investigated the roles of amylose and
from fine materials on both laboratory and commercial amylopectin from starch in the selective flocculation process and
scales (Dogu and Arol 2004). found that amylopectin is overwhelmingly responsible for the
The selective flocculation and dispersion process is the flocculant action of the starch. Amylopectin behaved almost as
only economically viable method of processing hematite ores a bulk flocculant, providing very high recoveries of hematite,
with iron grades less than 40% and liberation sizes smaller goethite, and kaolinite. Amylose suppressed the flocculation of
than 25 μm. However, it is only scarcely used in industry as all these minerals, but kaolinite more than the others. Thus, the
the reagent costs are relatively high and the process is very presence of amylose increased the overall quality of the separation
sensitive to surface chemistry changes (Haselhuhn, Carlson which could be achieved. The optimal amylose to amylopectin
and Kawatra 2012; Haselhuhn and Kawatra 2015c). ratio for selective flocculation was found to be around 1:3, which is
20 X. ZHANG ET AL.

Figure 14. Chemical names and structures of common iron ore dispersants (based on Haselhuhn 2015).

close to the natural composition of starch. Amylopectin’s floccula- a strongly hydrophobic configuration, precipitating out of
tion abilities are attributed to its high molecular weight and solution and forming hydrophobic aggregates. While poly-
branched molecular structure, while amylose’s ability to selectively acrylamide’s flocculation mechanism is irreversible,
suppress amylopectin’s flocculation is thought to stem from co- PNIPAM’s hydrophobic flocculation can easily be reverse
adsorption between the two molecules. by lowering the temperature. As such, PNIPAM can serve
Tripathy, Bhagat and Singh (2001) developed and investi- as the collector, flocculant, and dispersant. When compared
gated the flocculation characteristics of sodium alginate-g-poly- with sodium oleate, PNIPAM performs significantly better at
acrylamide grade VI (SAG-VI). When compared to some other concentrating hematite particles larger than 20 μm due to its
commercially available flocculants in aqueous suspensions of higher selectivity. PNIPAM is also found to be less sensitive
0.5wt% iron ore slimes, SAG-VI performed on par with or to the presence of fine materials.
somewhat better than the alternatives. It was concluded that Kemppainen et al. (2016) investigated the roles of two
iron ore slimes containing mostly smaller particle sizes are better types of anionic cellulose nanofibers when used in single-
flocculated by neutral flocculants than those with a larger parti- oxide mineral flocculation, particularly hematite and quartz.
cle size. Dicarboxylic acid (DCC) and sulfonated (ADAC) cellulose
Ng et al. (2015) proposed a novel approach for recovering nanofibers are found to selectively flocculate hematite in the
valuable fines using temperature-sensitive polymers. One pH range of 5–10, with little interaction observed between
such polymer is poly(N-isopropylacrylamide) (PNIPAM). the nanofibers and quartz. The tendency of DCC to form
Below approximately 32°C, the lower critical solution tem- strong flocs is found to depend on the pH of the suspension,
perature (LCST), PNIPAM is water soluble, hydrophilic, and with the largest quantity and most resilient flocs formed near
strongly dispersing. Above 32°C, PNIPAM transitions into pH 8–9 with a DCC dosage of 200–500ppm. DCC is capable
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW 21

of rapidly agglomerating hematite under vigorous stirring, gas flow rate, rate of agitation, the addition of wash water if any,
but ACAG requires longer and less rapid stirring to achieve and the physical properties of the particles (Dobby 2002; Pita
similar results. At 500ppm dosages however ACAG was 2015). However, the amount of entrainment can be strongly
found to be as effective as DCC while operating over correlated with the amount of water recovered in the froth, as
a wider pH range. shown in Figure 15. Since frothers control both the amount of
froth formed and the amount of water retained in the froth, an
excess of frother directly leads to an excessive amount of non-
4.3. Frothers selective recovery by entrainment. Two liquid flotation has been
used to determine whether material is being recovered by
It is common knowledge that frothers serve a prominent role in entrainment or true flotation (due to a failure of the depressant,
controlling the bubble size, froth stability, and froth mobility in for example) at two concentrators in Brazil (Nykänen et al. 2018).
flotation (Bulatovic 2007). For iron ore flotation, several different Suardini (1994) investigated controlling the rate of water
frothers are commonly used including pine oil, aliphatic alco- recovery by separating the frothing behavior of the collector
hols, propylene glycols, alkylethers of polypropylene, and cresylic from the frothing behavior of a non-collecting frother. These
acids. Most alcohols containing 5 to 8 carbon atoms can be results are shown in Figure 16, where it is observed that
utilized as frothers, including mixed amyl alcohol, methyl iso- a partial replacement of the etheramine of a cationic reverse
butyl carbinol, and sometimes heptanols or octanols (Nakhaei flotation circuit with methyl isobutyl carbinol (MIBC)
and Irannajad 2018). increased recovery by approximately 5 percentage points at
Although Houot (1983) reports that some specific alcohols a constant grade of 63.8%.
and propylene glycol have been used on occasion in the Araujo, Viana and Peres (2005) reports the investigation of
United States, they are rarely used in cationic reverse flotation partial replacement of an amine collector with some commer-
of hematite in the U.S. This is primarily because the cationic cial frothers at laboratory scale. These tests apparently pro-
flotation collectors, often ether amines, demonstrate strong vided promising results: the addition of a polyglycol frother
frothing properties at the pH ranges that cationic reverse could improve both the recovery and selectivity in several
flotation is performed at. This is sufficient for floating hema- tests. Pine oil frothers also improved performance, but linear
tite, but MIBC is often used during the flotation of magnetite and branched alcohols were less effective. Unfortunately, the
at a dosage of near 1% of the amine dosage, whether or not associated mineralogy is not directly reported.
the amine presents frothing properties. Rodrigues Silva, Correa De Araujo and Farias De Oliveira
Frothers are particularly important because there are two (2008) performed plant testing of this partial replacement
major mechanisms of recovery in flotation: true flotation, strategy, partially replacing the amine collector at the
which consists of air bubbles attaching to hydrophobized mate- Vargem Grande iron ore concentrator in Brazil with
rials and causing them to rise to the top; and entrainment, a frother. The frother used in this test was a blend of alcohols,
wherein solid particles become trapped between the rising bub- esters, and aldehydes, while the collector was a 20%-30%
bles. In the reverse flotation of iron minerals, the collector is used neutralized etheramine. The starch dosage, pulp pH, and
to hydrophobize the silica so that it can be floated. True flotation water chemistry were kept constant throughout these experi-
is highly selective, but entrainment is not (Nykänen et al. 2018). ments, while varying the quantity of collector replaced. These
Entrainment depends on many factors, including froth stability, trials found that a 10% replacement of the collector improved
the overall results: iron recovery was improved from 67.96%
to 80.78%, while iron grade was simultaneously improved
from 64.92% to 65.43%. This allowed for approximately an
extra $3 million/year in revenue. It was also found that less
neutralized etheramines allowed for a higher substitution of
frother before results declined.
Despite these significant results, it is still not clearly under-
stood as to which non-collecting frothers would provide the
best results. It may also be possible to further separate and
control the two reagent properties if a non-frothing collector
could be designed for cationic flotation.
Extensive research has been carried out to determine the
effect of frothers on bubble size and gas holdup in flotation.
Corona-Arroyo et al. (2015) characterized the impact of dode-
cylamine (DDA) along with methyl isobutyl carbinol (MIBC)
and polyglycol OREPREP™ F507. MIBC has a molecular
weight of 102 and a hydrophile-lipophile balance of 6.05,
Figure 15. Shown in the graph is a hypothetical recovery curve for reverse while F507 has a molecular weight of 425 and a hydrophile-
flotation. As the amount of water recovered increases, the amount of nonselec- lipophile balance of 8.63. The critical coalescence concentra-
tive recovery also increases. However, the rate at which silica is recovered grows tion (CCC), which is the concentration at which further
much faster, due to the presence of the collector. This means that the best
separations are achieved near the minimum possible water recoveries, and thus addition of reagent no longer decreases the bubble size, fol-
the lowest dosages of frothers (Kawatra and Eisele 2001). lowed the order of MIBC > DDA > F507 – that is to say that
22 X. ZHANG ET AL.

Figure 16. Batch reverse cationic flotation results wherein the etheramine collector was partially replaced with methyl isobutyl carbinol (Suardini 1994).

quantity of F507 required to achieve the maximum bubble movement of freely moving ions and polar compounds within
formation was the smallest of the three. While DDA is pri- the solution. However, some ions and compounds will remain
marily a collector, it is included in the comparison of frothers strongly adsorbed to the surface, even as the surface moves
due to its, in this case, non-negligible frothing properties. through the water. The boundary between the domain domi-
When the frothers (MIBC and F507) were mixed with the nated by the surface and the domain dominated by the bulk
DDA collector, the frother CCC and the resulting bubble sizes fluid is known as the shear plane. The effective surface charge
were observed to decrease. presented by the surface at the shear plane after accounting
Laskowski et al. (2003) reports that blending selective frothers for both the charge and volume of these attached ions is
(such as MIBC) with very small quantity of strong frothers (such known as the zeta potential. The pH at which the true surface
as polyglycol) result in froth behaviors more similar to the is neutral is known as point of zero charge (PZC), while the
stronger frother. That is, the small addition of polyglycol results pH at which the zeta potential is neutral is known as the
in smaller bubble sizes both below and above the CCC of MIBC. isoelectric point (IEP). In absence of specifically adsorbed
Tan et al. (2005) also found this behavior. Elmahdy and Finch ions or other chemisorbed species, the surface charge and
(2013) reported that this decrease bubble size only occurs below the zeta potential will tend to carry the same sign, and thus
the CCC of MIBC, and that above the CCC of MIBC the bubble the PZC and IEP are usually, though not necessarily, the same
size only decreases if the additional frother also reaches its CCC. value. Figure 17 presents the relation between the surface
charge and the presence of various sorts of adsorbing ions
with the zeta potential.
5. Principle surface chemical factors affecting iron
The surface charge and correspondingly the PZC are the pri-
ore flotation
mary factors determining the behavior of a surface in flotation. As
The selectivity of flotation depends on the complex interac- changing the pH changes the availability of hydronium cations
tion of mineralogical, chemical, physical, and engineering and hydroxide anions, the surface charge is subsequently shifted.
parameters. In an attempt to understand the key principles Figure 18 shows the typical behavior of silica and hematite surfaces
driving these interactions, a considerable amount of research under varying pH values. At high pH values, surfaces will tend to
effort has gone toward unraveling the effects of the surface become negatively charged, and at low pH values surfaces will tend
chemistry of iron ores and quartz. There have been several to become positively charged. The zeta potential again typically
extensive reviews of both the industrial practices and basic follows the sign of the surface charge, and largely determines the
principles of iron ore flotation (Houot 1983; Iwasaki 1983; interaction between two surfaces in the solution. Surfaces exhibit-
Numella and Iwasaki 1986). ing low zeta potentials are typically quick to flocculate, while
Most of the physics of iron ore flotation can be traced back to surfaces exhibiting large zeta potentials of the same sign will
interactions between the soluble ions in the process water and typically remain dispersed. Surfaces with opposing zeta potentials
the surfaces of the minerals. Understanding these interactions is may undergo heteroflocculation if the particles are fine enough,
critical to predicting and controlling the behavior of the surface but this is also the condition wherein simple direct flotation is most
chemical processes which lie at the heart of iron ore processing. applicable. After all, surfaces with negative charges can be selec-
tively collected with cationic collectors, and surfaces with positive
charges can be selectively collected with anionic collectors.
5.1. Effects of zeta potential
This interaction between surface charge and collector
Surfaces in water, except at the isoelectric point, tend to depends on the effective surface charge and the charge of
become charged by either dissociating protons at their surface the collector. At pH values below hematite’s PZC of 6.7
or adsorbing protons or other ions from the water (Uwadiale dodecylsulfonate is an effective collector, while above the
1992). These charges subsequently are balanced out by the PZC dodecylammonium is the effective collector (Iwasaki,
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW 23

Figure 17. The surface charge is measured at the inner Helmholtz plane, which includes only the effects of the surface and the specifically adsorbed ions. The zeta
potential is measured at the stern plane, which additionally contains the effects of other strongly adsorbed species, such as chemisorbed molecules or cations/anions
which move with the surface. Beyond the stern plane, dissolved ions tend to follow the liquid instead of the surface (based on Haselhuhn 2015).

Figure 18. Expected response of zeta potential to pH changes for hematite and silica. The value at which the zeta potential crosses 0 is known as the isoelectric
point, where the surface presents no net-charge.

Cooke and Colombo 1960; Uwadiale 1992). Note that while The PZC of quartz is pH 1.8. Below the PZC, anionic collec-
the PZC is typically a fundamental property of the surface, the tors are not typically adsorbed sufficiently to float quartz directly,
key to determining the adsorption of these chemicals is the as they must compete with the significant quantity of anions
IEP which may differ from the PZC in the presence of speci- already present in solution from the pH adjustment.
fically adsorbed ions such as Ca2+ in the case of iron ores. Additionally, at such acidic conditions some anionic collectors,
24 X. ZHANG ET AL.

particularly those based on weaker organic acids, may be forced presence of anionic collectors such as fatty acids (Clark 1968).
to collect the acidic protons in solution leaving them in an In cationic reverse flotation, Ca2+ plays apparently little role
uncharged form. Above the PZC, aminium (-RN+H3 groups) in the collection process, but Ca(OH)+ contributes to
readily adsorb to the quartz surface. Cases (1970) reports that a decrease in flotation recovery at pH >10.5. As previously
other silicates exhibit similar behavior. mentioned the adverse effects of Ca(OH)+ can be mitigated by
The surfaces of oxides and silicate minerals are often the use of an appropriate complexing agent such as STPP,
extensively hydrated, and can in turn form hydrogen bonds SHMP, or EDTA.
with the water phase. As such the adsorption of collectors, Magnesium also occurs in iron ore pulps. Under alkaline
and in particular the formation of collector hemimicelles, is conditions (pH >10), magnesium causes a sharp decrease in
often necessary to induce any form of hydrophobicity in these quartz recovery due to the precipitation of Mg(OH)2 which
compounds. coagulates onto the quartz surface. Mg(OH)2 is highly hydro-
Note that while pure hematite has a PZC of 6.7 (Uwadiale philic and thus does not float. This can also be remedied with
1992), natural hematite will often display vastly different IEPs. the use of complexing agents, but sodium silicate is ineffective
Very often, the IEP of natural hematite will appear to be at restoring the floatability of the afflicted quartz.
similar to the IEP of silica, which is attributed to the attach- The effect of magnesium on desliming and flotation can
ment of silica slimes to the hematite surface before desliming be far more pronounced than most other divalent cations.
or even before grinding (Carlson and Kawatra 2013). This is attributed to the small covalent diameter of the
It is also worth noting that the zeta potential plays a role in magnesium cation, 274 pm (Haselhuhn 2015; Pyykkö and
every aspect of iron ore refining, not just in flotation and Atsumi 2009). This is smaller than the spacing between
desliming. It can be helpful to be mindful of the roles that surface hydroxyl groups on hematite ore, which is approxi-
deslime and flotation reagents will play in later stages of the mately 285 pm (Haselhuhn 2015). This allows magnesium
beneficiation process. In particular, filtration is adversely to adsorb to and saturate every available surface site on
affected by strongly dispersing conditions, leading to slow hematite while calcium, which has a larger diameter of
filtration rates and which can be mitigated by reducing the 342 pm, can only adsorb to every other site (Haselhuhn
pH of filtration feed (Claremboux and Kawatra 2019). 2015; Pyykkö and Atsumi 2009).
Additionally, pelletization is adversely affected by strongly Iron cations, both Fe2+ and Fe3+, can be found in iron ore
flocculating conditions – while the addition of flocculants pulps processed in acidic conditions. The adsorption of these
makes pellet growth more rapid the strength of pellets cations onto magnetite surfaces can have a tremendous impact
decreases dramatically (Halt and Kawatra 2017). On the on the zeta potential (Sun et al. 1998). Even the addition of
other hand, dispersing conditions do not significantly slow 0.2 mM of iron (ii) chlorate increases the isoelectric point of
pellet growth rates, but lead to significantly stronger pellets magnetite from approximately 5 to over 7, and at alkaline pH
(Halt and Kawatra 2017). values causes the zeta potential to increase in magnitude from
approximately −20mV to −40mV (Sun et al. 1998). The addi-
tion of 0.02 mM iron (iii) chlorate does not increase the iso-
5.2. Effects of water chemistry
electric point as much, but does cause a comparable increase in
The importance of water quality on desliming and flotation of the magnitude of the zeta potential by a similar amount at all
iron ore has long been recognized. Fresh water may contain pH values above neutral (Sun et al. 1998). Interestingly, how-
any number of soluble ions. Most commonly fresh water ever, at 2 mM concentrations iron (iii) chlorate no longer shifts
contains some concentration of Na+, K+, Ca2+, and Mg2+. the zeta potential at alkaline pH values, but at acidic ones
Acidic fresh water may also contain Al3+, Fe2+, and Fe3+. instead (Sun et al. 1998). In general, larger concentrations of
Common anions in fresh water include HCO3−, SiO32-, and iron cations shift the IEP of the magnetite further toward
sometimes SO42- in locations where sulfides or gypsum occur. higher pH values (Sun et al. 1998).
Oxidized iron ores also typically contain some of these ions to The surface reactions controlling zeta potential may be
begin with, and may release these ions back into the water up represented by the equations below:
to concentrations of 0.1–1 mmol/L if the water is sufficiently
pure. These released ions may result in the accidental activa- ;FeðII; IIIÞOH þ Fe2þ or Fe3þ þ nOH
tion of silica in anionic reverse flotation and the accidental
¼ ;FeðII; IIIÞOFeðOHÞn1 ðn2 or 3Þ þ H2 O
depression of silica in cationic reverse flotation. Ca2+ may also
be intentionally added to activate silica in anionic reverse ðn increases with increasing pHÞ
flotation under highly alkaline conditions (Iwasaki 2000). Hþ Hþ
;FeðII; IIIÞOH2 þ ! ;FeðII; IIIÞOH ! ;FeðII; IIIÞO
The attachment of starches and clays to quartz or hematite
is largely determined by the water chemistry as many of these It is important also to realize that the surface adsorption of
ions are activators, such as calcium, or deactivators, such as metal ions to the minerals in iron ore can result in extre-
chloride in the context of starch adsorption on hematite mely different ion concentrations than in the bulk fluid.
(Lelis, Da Cruz and Lima 2019). Studies of an iron ore filter cake following filtration found
Iwasaki et al. (1980) reports that at neutral pH values that the filter cake moisture had an almost 25 times higher
calcium typically adsorbs as a Ca2+ cation, while at alkaline concentration of phosphate than the filter feed slurry did,
conditions (pH 11) it typically adsorbs as a Ca(OH)+ cation. and a 10 times higher concentration of sulfate (Haselhuhn
Ca(OH)+ is the necessary form for efficient activation in the and Kawatra 2015d). The effect on calcium can be even
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW 25

more dramatic, with one plant having a calcium concentra- flotation to remove silica. This avoids the difficulties
tion in the filter cake that was 565 times higher than in the of applying reverse flotation to carbonate rich iron
filtrate water (Eisele, Kawatra and Ripke 2005). The ten- ores, allowing for the practical utilization of
dency for ions to concentrate near the surfaces means that a massive reserve of previously stubborn iron ores.
even if the water is comparatively clean and deionized there (7) Desliming before flotation is an essential technology, as
may still be a significant concentration of trapped ions regardless of flotation route or reagents, an abundance of
within the ore itself, leading to changes in the zeta potential slimes will result in low selectivity, poor flotation rates,
and the activity of many reagents. and tremendous reagent usage. Anionic reverse flotation
is less sensitive than cationic reverse flotation, but is still
negatively impacted by the presence of large amounts of
6. Conclusions and recommendations
fine particles. Selective flocculation and dispersion
From the preceding review of the direct and reverse flota- remains the most common approach for desliming in
tion routes, desliming as a necessary precursor to many industry.
flotation routes, flotation reagents, and the primary sur- (8) The key to floating silica by reverse flotation is to
face chemistry factors affecting flotation, the following develop collector and depressant systems to effi-
conclusions and recommendations are provided: ciently separate iron ore from silicates. Further
research into high-efficiency and nontoxic cationic
(1) Flotation has been considered to be the most pro- collectors is recommended. If possible, finding addi-
mising method of obtaining useful, high-grade iron tional varieties and sources of cationic collectors
from low-grade, finely disseminated iron ores. would be preferable. Several novel cationic collectors
Flotation can be used alone or to complement such as M-302, EBAB, CS-22, TTPC, etc. have been
other separation methods, such as gravity separation tested successfully at laboratory scale, but remain
and magnetic separation. expensive and untested at industrial scales. Further
(2) Flotation is highly dependent on the surface chemistry research is recommended into starch alternatives.
of minerals to be separated. It is necessary to have PMTA, HA, and CMC may prove to be valuable
a firm understanding of effects of not only the starting points to developing more effective iron ore
reagents, but also the chemistry of the available plant depressants. The characteristic spacing within the
water and the impact of soluble ions in the ore. pyranose ring of amylose and amylopectin which
Selective flocculation and deslime is also very sensitive closely matches the surface spacing of hematite’s
to these parameters. iron atoms may be valuable when designing new
(3) Reverse flotation is the most promising flotation route depressants.
for further development. The choice of cationic versus (9) For anionic reverse flotation, the development of
anionic reverse flotation is dependent on mineralogy low-temperature anionic collectors is paramount.
and available reagents. Many facilities in China have The elevated temperature requirements of present
found that anionic reverse flotation is more suitable collectors result in complex dosing controls and
for their ores, but most of the rest of the world has high energy consumption by the flotation pro-
chosen cationic reverse flotation instead. cess. RA series collectors and MH-80 are cur-
(4) Cationic reverse flotation has the advantage of rently popular in industry, but laboratory results
higher flotation rates, simpler reagent systems, sim- for SO, α-BDA, etc. have also been promising.
ple and reliable operation, and can be operated at (10) Mixed collectors have shown considerable promise
low temperatures. However, cationic collectors have in improving flotation results while using fewer
relatively poor selectivity, foaming properties and reagents. Further investigation is recommended, as
higher toxicity, which can lead to product loss, lower reagent usage directly correlates to lower col-
highly viscous and poorly fluid froth, and environ- lector costs, lower production costs, and less poten-
mental pollution respectively. tial for environmental pollution.
(5) Anionic reverse flotation requires the use of an (11) For understanding flotation processes, the zeta
activator (usually lime) to float the quartz using an potential and the factors that go into determining
anionic collector at elevated pH values while depres- it remain paramount. Understanding the effects of
sing the iron with starch or another depressant. In process water chemistry and adsorbed ions on the
comparison to cationic reverse flotation, anionic zeta potential and the surface chemistry of miner-
reverse flotation is less sensitive to slimes, has als is one of the most important parts of under-
a lower collector cost, and high selectivity due to standing the theory of flotation. It is worth
the calcium activator having a combined activation repeating that the water chemistry can be con-
effect on the quartz and depression effect on the trolled by either adding salts or removing the
iron ore. However, reverse anionic flotation requires cations using complexing agents such as STPP,
a significant quantity of activator, along with higher SHMP, and EDTA.
process temperatures and alkalinities. (12) Researchers should continue to aim to connect
(6) Stepped flotation is a combination of direct flotation laboratory scale results with industrial scale results
of siderite from other iron ores followed by reverse to further optimize the flotation process.
26 X. ZHANG ET AL.

Disclosure statement Corona-Arroyo, M. A., A. López-Valdivieso, J. S. Laskowski, and


A. Encinas-Oropesa. 2015. Effect of frothers and dodecylamine on
No potential conflict of interest was reported by the authors. bubble size and gas holdup in a downflow column. Minerals
Engineering 81:109–15. doi:10.1016/j.mineng.2015.07.023.
DeVaney, F. D. 1985. Ore preparation and concentration methods. In
SME mineral processing Handbook, ed. N. L. Weiss. Section 20. New
References York: Society of Mining Engineers of the American Institute of
Mining, Metallurgical and Petroleum Engineers, Inc.
Abaka-Wood, G. B., J. Addai-Mensah, and W. Skinner. 2017. A study of Dobby, G. 2002. Column flotation. In Mineral processing plant design,
flotation characteristics of monazite, hematite, and quartz using anio- practice and control, ed. A. L. Mular, D. N. Halbe, and D. J. Barratt,
nic collectors. International Journal of Mineral Processing 158:55–62. 1239–52. Littleton: SME Inc.
doi:10.1016/j.minpro.2016.11.012. Dogu, I., and A. I. Arol. 2004. Separation of dark-colored minerals from
Araujo, A. C., P. R. M. Viana, and A. E. C. Peres. 2005. Reagents in iron feldspar by selective flocculation using starch. Powder Technology 139
ores flotation. Minerals Engineering 18 (2):219–24. doi:10.1016/j. (3):258–63. doi:10.1016/j.powtec.2003.11.009.
mineng.2004.08.023. Dos Santos, I. D., and J. F. Oliveira. 2007. Utilization of humic acid as
Arol, A. I., and I. Iwasaki. 1987. Control of montmorillonite via complexa- a depressant for hematite in the reverse flotation of iron ore. Minerals
tion and ultrasonics in the selective flocculation of iron ores. Mining, Engineering 20 (10):1003–07. doi:10.1016/j.mineng.2007.03.007.
Metallurgy & Exploration 4 (2):82–87. doi:10.1007/BF03403448. Edwards, C. R., W. B. Kipkie, and G. E. Agar. 1980. The effect of slime
Bruckard, W. J., L. K. Smith, and G. W. Heyes. 2015. Developments in coatings of the serpentine minerals, chrysotile and lizardite, on
the physiochemical separation of iron ore. In Iron ore, ed. L. Lu, pentlandite flotation. International Journal of Mineral Processing 7
339–56. UK: Elsevier. (1):33–42. doi:10.1016/0301-7516(80)90035-6.
Bulatovic, S. M. 2007. Flotation reagents: Chemistry, theory and practice, Eisele, T. C., S. K. Kawatra, and S. J. Ripke. 2005. Water chemistry effects in
3–99. Netherlands: Elsevier Science. iron ore concentrate agglomeration feed. Mineral Processing & Extractive
Bunge, F. H., J. B. Morrow, and L. W. Trainor. 1977. Developments and Metallurgy Review 26:295–305. doi:10.1080/08827500590944063.
realizations in the flotation of iron ores in North America. Iron, XIIth Elmahdy, A. M., and J. A. Finch. 2013. Effect of frother blends on
IMPC, Sao Paulo. hydrodynamic properties. International Journal of Mineral Processing
Cao, Z., Y. Zhang, and Y. Cao. 2013. Reverse flotation of quartz from 123:60–63. doi:10.1016/j.minpro.2013.04.019.
magnetite ore with modified sodium oleate. Mineral Processing and Fan, Y., L. Jing, and S. Zhu. 2011. Research and application status of
Extractive Metallurgy Review 34 (5):320–30. doi:10.1080/08827508. reverse flotation collectors for refractory iron ore. Shandong
2012.675531. Metallurgy 6:6. (in Chinese).
Carlson, J. J., and S. K. Kawatra. 2013. Factors affecting zeta potential of Filippov, L. O., I. V. Filippova, and V. V. Severov. 2010. The use of
iron oxides. Mineral Processing and Extractive Metallurgy Review 34 collectors mixture in the reverse cationic flotation of magnetite ore:
(5):269–303. doi:10.1080/08827508.2011.604697. The role of Fe-bearing silicates. Minerals Engineering 23 (2):91–98.
Cases, J. M. 1970. On the normal interaction between adsorbed species doi:10.1016/j.mineng.2009.10.007.
and adsorbing surface. Transactions of AIME 247 (2):123. Filippov, L. O., V. V. Severov, and I. V. Filippova. 2014. An overview of
Chen, D., Y. Ge, and Y. Yu. 2005. The study of reverse flotation the beneficiation of iron ores via reverse cationic flotation.
technology for iron concentrate of magnetic separation and its International Journal of Mineral Processing 127:62–69. doi:10.1016/j.
reagents. Conservation and Utilization of Mineral Resources 4:46– minpro.2014.01.002.
50. (in Chinese). Forsmo, S. P. E., S. E. Forsmo, B. M. T. Björkman, and P. O. Samskog.
Chen, P., Y. Hu, Z. Gao, J. Zhai, D. Fang, T. Yue, and W. Sun. 2017. Discovery 2008. Studies on the influence of a flotation collector reagent on iron
of a novel cationic surfactant: Tributyltetradecyl-phosphonium chloride ore green pellet properties. Powder Technology 182:444–52.
for iron ore flotation: From prediction to experimental verification. doi:10.1016/j.powtec.2007.07.015.
Minerals 7 (12):240. doi:10.3390/min7120240. Fuerstenau, D. W., A. M. Gaudin, and H. L. Miaw. 1958. Iron oxide
Chen, X., and Y. Peng. 2018. Managing clay minerals in froth flotation – slime coatings in flotation. Trans. AIME 211:792–93.
A critical review. Mineral Processing and Extractive Metallurgy Review Fuerstenau, M. C., and B. R. Palmer. 1976. Anionic flotation of oxides and
39 (5):289–307. doi:10.1080/08827508.2018.1433175. silicates. In Flotation-A. M. Gaudin memorial, ed. M.C. Fuerstenau, 148.
Chi, F. H., and G. L. Amy. 2004. Kinetic study on the sorption of New York: AIME.
dissolved natural organic matter onto different aquifer materials: Fuerstenau, M. C., J. Jameson, and R. Yoon. 2007. Froth flotation:
The effects of hydrophobicity and functional groups. Journal of A century of innovation. Littleton, CO: SME.
Colloid and Interface Science 274 (2):380–91. doi:10.1016/j. Gagnon, B. 1981. Iron ore company of Canada Zv concentrator flowsheet
jcis.2003.12.049. description and process development. Superintendent Sept-Iles.
Claremboux, V., and S. K. Kawatra. 2019. Application of surface Ge, Y., J. Yu, and P. Zhu. 2009. Review of flotation reagents for iron ore.
chemical fundamentals to improving industrial filtration rates. Modern Mining 11:6–11. (in Chinese).
Mineral Processing & Extractive Metallurgy Review 40 (4):292–97. Halt, J. A., and S. K. Kawatra. 2017. Does the zeta potential of an iron ore
doi:10.1080/08827508.2019.1598404. concentrate affect the strength and dustiness of unfired and fired
Clark, S. W. 1968. Adsorption of calcium, magnesium, and sodium ion pellets? Mineral Processing & Extractive Metallurgy Review 38
by quartz. Transactions of the Society of Mining Engineers of AIME 241 (2):132–41. doi:10.1080/08827508.2017.1288114.
(3):334–41. Han, Y., Z. Yuan, Y. Li, and B. Chen. 2006. Advances in mineral
Clemmer, J. B. 1947. Flotation of iron ore. 8th Annual Mining processing technology of China metallic mine and its development
Symposium, Duluth, Minnesota, U.S.A. orientation. Metal Mine 1:34–52. (in Chinese).
Colombo, A. F., and D. W. Frommer. 1976. Cationic flotation of mesabi Haselhuhn, H. J. 2015. The dispersion and selective flocculation of
range oxidized taconite. In Flotation A.M. Gaudin memorial volume, hematite ore. Dissertation, Michigan Technological University.
ed. M. C. Fuerstenau, 1285–304. USA: SME-AIME. Haselhuhn, H. J., J. J. Carlson, and S. K. Kawatra. 2012. Water chemistry
Colombo, A. F. 1980. Selective flocculation and flotation of iron bearing analysis of an industrial selective flocculation dispersion hematite ore
materials. In Fine particles processing, ed. P. Somasundaran, Vol. 2, concentrator plant. International Journal of Mineral Processing
1034–56. New York: SME-AIME. 102:99–106. doi:10.1016/j.minpro.2011.10.002.
Colombo, A. F. 1986. Concentration of iron oxides by selective Haselhuhn, H. J., and S. K. Kawatra. 2015a. Effects of water chemistry
flocculation-flotation. In Advances in Mineral Processing: A Half- on hematite selective flocculation and dispersion. Mineral Processing
Century of Progress in Application of Theory to Practice, 695–713. and Extractive Metallurgy Review 36 (5):305–09. doi:10.1080/
Englewood, CO: Society of Mining Engineers of AIME. 08827508.2014.978318.
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW 27

Haselhuhn, H. J., and S. K. Kawatra. 2015b. Flocculation and dispersion Chemical Engineering Science 148:256–66. doi:10.1016/j.
studies of iron ore using laser scattering particle size analysis. Minerals ces.2016.04.014.
& Metallurgical Processing 32 (4):191–95. Krishnan, S. V., and I. Iwasaki. 1984. Pulp dispersion in selective
Haselhuhn, H. J., and S. K. Kawatra. 2015c. Role of water chemistry in desliming of iron ores. International Journal of Mineral Processing
the selective flocculation and dispersion of iron ore. Mining, 12 (1–3):1–13. doi:10.1016/0301-7516(84)90019-X.
Metallurgy & Exploration 32 (2):69–77. doi:10.1007/BF03402423. Kruger, M., S. Naik, and N. Naudé. 2018. Upgrade of SLon concentrate
Haselhuhn, H. J., and S. K. Kawatra. 2015d. Aqueous ions in process with the use of froth flotation on a typical African iron ore. Mineral
water and cake moisture during iron ore filtration. Mineral Processing Processing and Extractive Metallurgy Review 39 (3):152–66.
and Extractive Metallurgy Review 36 (6):370–76. doi:10.1080/ doi:10.1080/08827508.2017.1399886.
08827508.2015.1004401. Kumar, T. V. V., D. S. Rao, S. Subba Rao, S. Prabhakar, and G. Bhaskar
Heerema, R. H., R. J. Lipp, and I. Iwasaki. 1982. Complexation of calcium Raju. 2010. Reverse flotation studies on an Indian low grade iron ore
ion in selective flocculation of iron ores. Trans. SME-AIME 272:1879–84. slimes. International Journal of Engineering Science and Technology 2
Holland, P. M., and D. N. Rubingh. 1992. Mixed surfactant systems, Vol. (4):637–48.
1. Washington, DC: American Chemical Society. Lai, C. H., and C. Y. Chen. 2001. Removal of metal ions and humic acid
Houot, R. 1983. Beneficiation of iron ore by flotation-review of industrial from water by iron-coated filter media. Chemosphere 44 (5):1177–84.
and potential applications. International Journal of Mineral Processing doi:10.1016/S0045-6535(00)00307-6.
10 (3):183–204. doi:10.1016/0301-7516(83)90010-8. Laskowski, J. S., T. Tlhone, P. Williams, and K. Ding. 2003. Fundamental
Hu, L., H. Zheng, and G. Xiao. 2010. Development of iron ore flotation properties of the polyoxypropylene alkyl ether flotation frothers.
technology. Modern Mining 1:23–26. (in Chinese). International Journal of Mineral Processing 72 (1–4):289–99.
Huang, Y., G. Han, J. Liu, and W. Wang. 2016. A facile disposal of Bayer doi:10.1016/S0301-7516(03)00105-4.
red mud based on selective flocculation desliming with organic Leal Filho, L. S., and G. A. Rodrigues. 1992. The use of ethoxylated
humics. Journal of Hazardous Materials 301:46–55. doi:10.1016/j. nonionic surfactants on the cationic flotation of quartz. Flotation:
jhazmat.2015.08.035. Fundamentals, Practice and Environment. Proceedings of the III
Huang, Z., H. Zhong, S. Wang, L. Xia, W. Zou, and G. Liu. 2014. Meeting of the Southern Hemisphere on Mineral Technology,
Investigations on reverse cationic flotation of iron ore by using ABTM, Belo Horizonte, Brazil, 50–64.
a Gemini surfactant: Ethane-1, 2-bis (dimethyl-dodecyl-ammonium Lelis, D. F., D. G. Da Cruz, and R. M. F. Lima. 2019. Effects of calcium
bromide). Chemical Engineering Journal 257:218–28. doi:10.1016/j. and chloride ions in iron ore reverse cationic flotation: Fundamental
cej.2014.07.057. studies. Mineral Processing & Extractive Metallurgy Review 40:402–09.
Illés, E., and E. Tombácz. 2003. The role of variable surface charge and doi:10.1080/08827508.2019.1666122.
surface complexation in the adsorption of humic acid on magnetite. Lima, N. P. 2001 Behavior of itabirite ores from Alegria and Fábrica Nova
Colloids and Surfaces A: Physicochemical and Engineering Aspects 230 deposits in the desliming and flotation processes. M.Sc thesis, Curso de
(1–3):99–109. doi:10.1016/j.colsurfa.2003.09.017. Pós Graduação em Engenharia Metalúrgica e de Minas, Universidade
Iwasaki, I. 1983. Iron ore flotation, theory and practice. Mining Engineering Federal de Minas Gerais. March de 2001. (In Portuguese).
35:622–31. Lima, N. P., A. E. C. Peres, and M. L. S. Marques. 2012. Effect of slimes
Iwasaki, I. 1999. Iron ore flotation-historical perspective and future on iron ores flotation. International Journal of Mining Engineering and
prospects. Advances in Flotation Technology as held at the Mineral Processing 1 (2):43–46. doi:10.5923/j.mining.20120102.04.
1999 SME Annual Meeting, Denver, CO, 231–43. doi:10.1016/ Lima, N. P., G. E. S. Valadão, and A. E. C. Peres. 2013. Effect of amine
s0378-4347(99)00464-8. and starch dosages on the reverse cationic flotation of an iron ore.
Iwasaki, I. 2000. Iron ore beneficiation in the USA: Past and future, 4–5. Minerals Engineering 45:180–84. doi:10.1016/j.mineng.2013.03.001.
USA: University of Minnesota. Liu, D. 2003. Current development state and prospect of the reverse
Iwasaki, I., K. A. Smith, R. J. Lipp, and H. Sato. 1980. Effect of calcium flotation process used in increasing iron grade and decreasing silicon
and magnesium ions on selective desliming and cationic flotation of content in iron concentrates. Metal Mine 2:38–42. (in Chinese).
quartz from iron ores. Fine Part Process, Proceedings of the Liu, J., J. Zhang, and J. Liu. 2007. Status quo of iron ores flotation
International Symposium, AIME, New York. reagent. China Mining Magazine 16 (2):106–08. (in Chinese).
Iwasaki, I., and R. W. Lai. 1965. Starches and starch products as depres- Liu, M., and Y. Shi. 2009. Study on the selective flocculation-desliming
sants in soap flotation of activated silica from iron ores. Trans. AIME and cationic reverse flotation technique for silicate removal from an
232:364–71. iron ore. Mining and Metallurgical Engineering 2:29–31. (in Chinese).
Iwasaki, I., S. R. B. Cooke, and A. F. Colombo. 1960. Flotation character- Liu, W., W. Liu, B. Wang, H. Duan, X. Peng, X. Chen, and Q. Zhao. 2019a.
istics of goethite, Vol. 5593, 24–25. RI: U.S. Bur. Mines. Novel hydroxy polyamine surfactant N-(2-hydroxyethyl)-N-dodecyl-
Iwasaki, I., S. R. B. Cooke, D. H. Harraway, and H. S. Choi. 1962a. Iron ethanediamine: Its synthesis and flotation performance study to quartz.
wash ore slimes-some mineralogical and flotation characteristics. Minerals Engineering 142:105894. doi:10.1016/j.mineng.2019.105894.
Trans. AIME 223:97–108. Liu, W., W. Liu, B. Zhao, L. Zhao, D. Li, P. Fang, and W. Liu. 2019b.
Jain, V., B. Rai, U. V. Waghmare, V. Tammishetti, and Pradip. 2013. Novel insights into the adsorption mechanism of the isopropanol
Processing of alumina-rich iron ore slimes: Is the selective dispersion– amine collector on magnesite ore: A combined experimental and
flocculation–flotation the solution we are looking for the challenging theoretical computational study. Powder Technology 343:366–74.
problem facing the Indian iron and steel industry? Indian Institute of doi:10.1016/j.powtec.2018.11.063.
Metals 66:447–56. doi:10.1007/s12666-013-0287-1. Luo, B., Y. Zhu, C. Sun, Y. Li, and Y. Han. 2015. Flotation and
Jena, S. K., H. Sahoo, S. S. Rath, D. S. Rao, S. K. Das, and B. Das. 2015. adsorption of a new collector α-Bromodecanoic acid on quartz sur-
Characterization and processing of iron ore slimes for recovery of iron face. Minerals Engineering 77:86–92. doi:10.1016/j.
values. Mineral Processing and Extractive Metallurgy Review 36 mineng.2015.03.003.
(3):174–82. doi:10.1080/08827508.2014.898300. Luo, X. M., W. Z. Yin, J. Yao, C. Y. Sun, Y. Cao, Y. Q. Ma, and Y. Hou.
Kar, B., H. Sahoo, S. S. Rath, and B. Das. 2013. Investigations on different 2013. Flotation separation of magnetic separation concentrate of
starches as depressants for iron ore flotation. Minerals Engineering refractory hematite containing carbonate with enhanced dispersion.
49:1–6. doi:10.1016/j.mineng.2013.05.004. Chinese Journal of Nonferrous Metals 23:238. (in Chinese).
Kawatra, S. K., and T. C. Eisele. 2001. Coal desulfurization: High- Ma, M. 2012. Froth flotation of iron ores. International Journal of Mining
efficiency preparation methods, 140. New York: Taylor & Francis. Engineering and Mineral Processing 1 (2):56–61. doi:10.5923/j.
Kemppainen, K., T. Suopajärvi, O. Laitinen, A. Ämmälä, mining.20120102.06.
H. Liimatainen, and M. Illikainen. 2016. Flocculation of fine hema- Ma, X. 2008. Role of solvation energy in starch adsorption on oxide
tite and quartz suspensions with anionic cellulose nanofibers. surfaces. Colloids and Surfaces A: Physicochemical and Engineering
Aspects 320:36–42. doi:10.1016/j.colsurfa.2008.01.011.
28 X. ZHANG ET AL.

Ma, X. 2010. Role of hydrolyzable metal cations in starch-kaolinite Peres, A. E. C., N. P. Lima, and A. C. Araújo. 2003. How different iron
interactions. International Journal of Mineral Processing 97:100–03. ore types behave in desliming in hydrocyclones and flotation.
doi:10.1016/j.minpro.2010.09.003. Hydrocyclones 3:1–6.
Ma, X. 2011 The dispersion of kaolinite. Iron Ore Conference, Perth, Pinto, C. L. L., A. C. Araujo, and A. E. C. Peres. 1992. The effect of starch,
Australia, July, 471–74. doi:10.1016/j.atherosclerosis.2011.02.032. amylose and amylopectin on the depression of oxi-minerals. Minerals
Ma, X., M. Marques, and C. Gontijo. 2011. Comparative studies of reverse Engineering 5 (3–5):469–78. doi:10.1016/0892-6875(92)90226-Y.
cationic/anionic flotation of vale iron ore. International Journal of Pita, F. A. 2015. True flotation and entrainment of kaolinitic ore in batch
Mineral Processing 100 (3–4):179–83. doi:10.1016/j.minpro.2011.07.001. tests. Mineral Processing & Extractive Metallurgy Review 36:213–25.
Ma, X., and W. J. Bruckard. 2010. The effect of pH and ionic strength on doi:10.1080/08827508.2014.928619.
starch-kaolinite interactions. International Journal of Mineral Poperechnikova, O. Y., L. O. Filippov, E. N. Shumskaya, and I. V. Filippova.
Processing 94:111–14. doi:10.1016/j.minpro.2010.01.004. 2017. Intensification of the reverse cationic flotation of hematite ores
Ma, X., W. J. Bruckard, and R. Holmes. 2009. Effect of collector, pH and with optimization of process and hydrodynamic parameters of flotation
ionic strength on the cationic flotation of kaolinite. International Journal cell. Journal of Physics: Conference Series 879:1–8.
of Mineral Processing 93:54–58. doi:10.1016/j.minpro.2009.05.007. Pradip, I. 1994. Reagents design and molecular recognition at mineral
Manukonda, V. R., and I. Iwasaki. 1987. Control of calcium ion via surfaces. In Reagents for Better Metallurgy, ed. P.S. Mulukutla, 245–52.
chemical precipitation-ultrasonic treatment in selective flocculation. Littleton, CO: Society for Mining, Metallurgy, and Exploration.
Mining, Metallurgy & Exploration 4 (4):217–22. doi:10.1007/ Pyykkö, P., and M. Atsumi. 2009. Molecular single-bond covalent radii
BF03402696. for elements 1-118. Chemistry – A European Journal 15 (1):186–97.
Mao, F., L. Huang, and F. Zhao. 2005. Achievements and development doi:10.1002/chem.v15:1.
prospect of mineral processing technology of China’s iron mines. Quast, K. 2006. Flotation of hematite using C6–C18 saturated fatty acids.
Metal Mine 2:1. (in Chinese). Minerals Engineering 19 (6–8):582–97. doi:10.1016/j.mineng.2005.09.010.
McDonald, J. E., and S. K. Kawatra. 2017. Agglomeration of hematite Ramos-Tejada, M. M., A. Ontiveros, J. L. Viota, and J. D. G. Durán.
concentrate by starches. Mineral Processing and Extractive Metallurgy 2003. Interfacial and rheological properties of humic acid/hematite
Review 38 (1):1–6. doi:10.1080/08827508.2016.1233875. suspensions. Journal of Colloid and Interface Science 268 (1):85–95.
Mohanty, S., and B. Das. 2010. Optimization studies of hydrocylone doi:10.1016/S0021-9797(03)00665-9.
for beneficiation of iron ore slimes. Mineral Processing & Rao, K., R. K. Dwari, S. Lu, A. Vilinska, and P. Somasundaran. 2011.
Extractive Metallurgy Review 31 (2):86–96. doi:10.1080/ Mixed anionic/non-ionic collectors in phosphate gangue flotation
08827500903397142. from magnetite fines. Open Mineral Processing Journal 4:14–24.
Nakhaei, F., and M. Irannajad. 2018. Reagents types in flotation of iron doi:10.2174/18748414001104010014.
oxide minerals: A review. Mineral Processing and Extractive Rao, K. H., B. M. Antti, and K. S. E. Forssberg. 1990. Flotation of mica
Metallurgy Review 39 (2):89–124. doi:10.1080/08827508.2017.1391245. minerals and selectivity between muscovite and biotite while using
Ng, W. S., R. Sonsie, E. Forbes, and G. V. Franks. 2015. Flocculation/ mixed anionic/cationic collectors. Mining, Metallurgy & Exploration 7
flotation of hematite fines with anionic temperature-responsive poly- (3):127–32. doi:10.1007/BF03403286.
mer acting as a selective flocculant and collector. Minerals Engineering Rao, K. H., and E. Forssberg. 1993. Solution chemistry of mixed cationic/
77:64–71. doi:10.1016/j.mineng.2015.02.013. anionic collectors and flotation separation of feldspar from quartz.
Numella, W., and I. Iwasaki, 1986. Iron ore flotation. In Advances in Mineral International Mineral Processing Congress: 23/05/1993-28/05/1993,
Processing: A Helf Century of Progress in Application of Theory to Practice, Sydney, Australia, Vol. 4, 837–44.
308–42. Englewood, CO: Society of Mining Engineers of AIME Rao, K. H., and K. S. E. Forssberg. 1997. Mixed collector systems in
Nykänen, V. P. S., A. S. Braga, T. C. S. Pinto, P. H. Matai, N. P. Lima, flotation. International Journal of Mineral Processing 51 (1–4):67–79.
L. S. Leal Filho, and M. B. Monte. 2018. True flotation versus entrain- doi:10.1016/S0301-7516(97)00039-2.
ment in reverse cationic flotation for the concentration of iron ore at Rao, S. 2004. Surface chemistry of froth flotation. In Second edition, Rev.
industrial scale. Mineral Processing and Extractive Metallurgy Review edn. of: Surface chemistry of froth flotation/Jan Leja, 32–34. New York:
1–11. doi:10.1080/08827508.2018.1514298. Kluwer Academic/Plenum Publishers.
Oyarzun, F. C. 2013. Modificaciones del proceso de flotación inversa de Rao, S. R. 2013. Surface chemistry of froth flotation: Volume 1:
hierro en celdas neumáticas de Planta Magnetita. Valparaíso, Chile: Fundamentals. New York: Springer Science & Business Media.
Pontificia Universidad Católica de Valparaíso, Facultad de Ingeniería. Ravishankar, S. A., I. Pradip, and N. K. Khosla. 1995. Selective flocculation of an
(In Spanish). iron oxide from its synthetic mixtures with clays: A comparison of poly-
Panda, L., P. K. Banerjee, S. K. Biswal, R. Venugopal, and N. R. Mandre. acrylic acid and starch polymers. International Journal of Mineral Processing
2013. Performance evaluation for selectivity of the flocculant on 43 (3–4):235–47. doi:10.1016/0301-7516(95)00011-2.
hematite in selective flocculation. International Journal of Minerals, Rodrigues Silva, R. R., A. Correa De Araujo, and J. Farias De Oliveira.
Metallurgy, and Materials 20 (12):1123–29. doi:10.1007/s12613-013- 2008. Frother assisted amine flotation of iron ores. 2nd International
0844-y. Iron Symposium, São Luís, Maranhão, Brazil.
Papini, R. M., P. R. G. Brandao, and A. E. C. Peres. 2001. Cationic flotation Sahoo, H., S. S. Rath, B. Das, and B. K. Mishra. 2016. Flotation of quartz using
of iron ores: Amine characterization and performance. Mining, ionic liquid collectors with different functional groups and varying chain
Metallurgy & Exploration 18 (1):5–9. doi:10.1007/BF03402863. lengths. Minerals Engineering 95:107–12. doi:10.1016/j.mineng.2016.06.024.
Pavlovic, S., and P. R. G. Brandao. 2003. Adsorption of starch, amylose, Sahoo, H., S. S. Rath, S. K. Jena, B. K. Mishra, and B. Das. 2015. Aliquat-
amylopectin and glucose monomer and their effect on the flotation of 336 as a novel collector for quartz flotation. Advanced Powder
hematite and quartz. Minerals Engineering 16 (11):1117–22. Technology 26 (2):511–18. doi:10.1016/j.apt.2014.12.010.
doi:10.1016/j.mineng.2003.06.011. Sandvik, K. L., and E. Larsen. 2014. Iron ore flotation with environmen-
Pereira, S. R. N., A. E. C. Peres, A. C. Araujo, and G. E. S. Valadão 2006. tally friendly reagents. Mining, Metallurgy & Exploration 31
The use of non-polar oil in the reverse cationic flotation of an iron (2):95–102. doi:10.1007/BF03402418.
ore. Proceedings of the XXIII International Mineral Processing Shao, A. 2013. Flotation separation of Donganshan carbonates-containing
Congress, Istambul, Turkey. hematite ore. Journal of Central South University: Science and Technology
Peres, A. E. C., A. C. Araujo, H. El-Shall, P. Zhang, and N. A. Abdel- 2:456–60. (in Chinese).
Khalek. 2009. Plant practice: Nonsulfide minerals. In Froth flotation: Sis, H., and S. Chander. 2003. Reagents used in the flotation of phos-
A century of innovation, Ch. 5, ed. M.C. Fuerstenau, G.J. Jameson, and phate ores: A critical review. Minerals Engineering 16 (7):577–85.
R.-H. Yoon, 857–61. Colorado: SME. doi:10.1016/S0892-6875(03)00131-6.
Peres, A. E. C., and M. I. Correa. 1996. Depression of iron oxides with Song, B., L. Yuan, and S. Wei. 2015. Investigation on stepped flotation
corn starches. Minerals Engineering 9 (12):1227–34. doi:10.1016/ process for carbonate-containing hematite and production practice.
S0892-6875(96)00118-5. Mining, Metallurgical Engineers 35:63–67. (in Chinese).
MINERAL PROCESSING AND EXTRACTIVE METALLURGY REVIEW 29

Song, R., and W. Li. 2008. Review of the flotation technology develop- Wang, X. 2009. Effect of desiliming on flotation of magnetite. Morden
ment of Anshan steel’s mines. Metal Mine 9:1–6. (in Chinese). Mining 486:117–19. (in Chinese).
Suardini, P. K. 1994. Analysis of the behavior of iron ore and coal flotation Wang, Y., and J. Ren. 2005. The flotation of quartz from iron
circuits. Thesis for the Degree of M.S., Michigan Technological minerals with a combined quaternary ammonium salt.
University, 243. doi:10.3168/jds.S0022-0302(94)77044-2. International Journal of Mineral Processing 77 (2):116–22.
Sun, B. 2006. Progress in China’s beneficiation technology for complex doi:10.1016/j.minpro.2005.03.001.
refractory iron ore. Metal Mine 3:11–13. (in Chinese). Weisseborn, P. K., L. J. Warren, and J. G. Dunn. 1995. Selective floccula-
Sun, Z. X., F. W. Su, W. Forsling, and P. O. Samskog. 1998. Surface character- tion of ultrafine iron ore. 1. Mechanism of adsorption of starch onto
istics of magnetite in aqueous suspension. Journal of Colloid and Interface hematite. Colloids and Surfaces A: Physicochemical and Engineering
Science 197 (1):151–59. doi:10.1006/jcis.1997.5239. Aspects 99 (1):11–27. doi:10.1016/0927-7757(95)03111-P.
Takahashi, Y., T. Kumano, and S. Nishikawa. 2004. Crystal structure of Weissenborn, P. K. 1996. Behaviour of amylopectin and amylose com-
B-Amylose. Macromolecules 37:6827–32. doi:10.1021/ma0490956. ponents of starch in the selective flocculation of ultrafine iron ore.
Tan, S. N., R. J. Pugh, D. Fornasiero, R. Sedev, and J. Ralston. 2005. Foaming International Journal of Mineral Processing 47 (3–4):197–211.
of polypropylene glycols and glycol/MIBC mixtures. Minerals doi:10.1016/0301-7516(95)00096-8.
Engineering 18 (2):179–88. doi:10.1016/j.mineng.2004.08.017. Weissenborn, P. K., L. J. Warren, and J. G. Dunn. 1994. Optimisation of
Tang, Z., P. Gao, Y. Sun, Y. Han, E. Li, J. Chen, and Y. Zhang. 2019. selective flocculation of ultrafine iron ore. International Journal of Mineral
Studies on the fluidization performance of a novel fluidized bed Processing 42 (3–4):191–213. doi:10.1016/0301-7516(94)00026-3.
reactor for iron ore suspension roasting. Powder Technology. Weng, X., G. Mei, T. Zhao, and Y. Zhu. 2013. Utilization of novel
doi:10.1016/j.powtec.2019.09.092. ester-containing quaternary ammonium surfactant as cationic collec-
Terashima, M., M. Fukushima, and S. Tanaka. 2004. Influence of pH tor for iron ore flotation. Separation and Purification Technology
on the surface activity of humic acid: Micelle-like aggregate forma- 103:187–94. doi:10.1016/j.seppur.2012.10.015.
tion and interfacial adsorption. Colloids and Surfaces A: Yang, Z., Q. Teng, J. Liu, W. Yang, D. Hu, and S. Liu. 2019. Use of
Physicochemical and Engineering Aspects 247 (1–3):77–83. NaOL and CTAB mixture as collector in selective flotation separa-
doi:10.1016/j.colsurfa.2004.08.028. tion of enstatite and magnetite. Colloids and Surfaces A:
Thella, J. S., A. K. Mukherjee, and N. G. Srikakulapu. 2012. Processing of Physicochemical and Engineering Aspects 570 (5):481–86.
high alumina iron ore slimes using classification and flotation. Powder doi:10.1016/j.colsurfa.2019.03.064.
Technology 217:418–26. doi:10.1016/j.powtec.2011.10.058. Yin, W., Y. Ma, M. Liu, M. A. Traore, M. Zhang, and L. Li. 2011.
Tian, J., L. Xu, Y. Yang, J. Liu, X. Zeng, and W. Deng. 2017. Selective Industrial tests on step-flotation of iron ore containing high ferric
flotation separation of ilmenite from titanaugite using mixed anionic/ carbonate in Donganshan. Metal Mine 40 (8):64–67. (in Chinese).
cationic collectors. International Journal of Mineral Processing Yin, W. Z., Y. X. Han, and F. Xie. 2010. Two-step flotation recovery of
166:102–07. doi:10.1016/j.minpro.2017.07.006. iron concentrate from Donganshan carbonaceous iron ore. Journal of
Tombacz, E., Z. Libor, E. Illes, A. Majzik, and E. Klumpp. 2004. The Central South University of Technology 17 (4):750–54. (in Chinese).
role of reactive surface sites and complexation by humic acids in doi:10.1007/s11771-010-0551-z.
the interaction of clay mineral and iron oxide particles. Organic Yong, Y. 2005. Processing state and technology progress of iron ore in China.
Geochemistry 35 (3):257–67. doi:10.1016/j.orggeochem.2003.11.002. Conservation and Utilization of Mineral Resources 6:102–06. (in Chinese).
Tripathy, S. K., P. K. Banerjee, and N. Suresh. 2015. Effect of desliming Yu, J., Y. Han, Y. Li, and P. Gao. 2019. Recent advances in magnetization
on the magnetic separation of low-grade ferruginous manganese ore. roasting of refractory iron ores: A technological review in the past
International Journal of Minerals, Metallurgy, and Materials 22 decade. Mineral Processing and Extractive Metallurgy Review 1–11.
(7):661–73. doi:10.1007/s12613-015-1120-0. doi:10.1080/08827508.2019.1634565.
Tripathy, T., R. P. Bhagat, and R. P. Singh. 2001. The flocculation Yuan, Z., Y. Han, and W. Yin. 2007. Status quo and development orientation of
performance of grafted sodium alginate and other polymeric floc- China’s refractory ore resource utilization. Metal Mine 1:01. (in Chinese).
culants in relation to iron ore slime suspension. European Polymer Zhang, C., and H. Dai. 2012. Current situation and future development
Journal 37 (1):125–30. doi:10.1016/S0014-3057(00)00089-6. of the collector for reverse flotation of iron ore. Multipurpose
Turrer, H. D. G., and A. E. C. Peres. 2010. Investigation on alternative Utilization of Mineral Resources 2:3–6. (in Chinese).
depressants for iron ore flotation. Minerals Engineering 23 (11–- Zhang, M., M. Liu, W. Yin, Y. Han, and Y. Li. 2007. Investigation on
13):1066–69. doi:10.1016/j.mineng.2010.05.009. stepped-flotation process for Donganshan carbonate-containing
USGS, 2017, "Mineral commodity summaries 2017." U.S. Geological Survey, 91. refrac-tory iron ore. Metal Mine 9:62–64. (in Chinese).
Usui, S. 1972. Heterocoagulation. In Progress in Surface and Membrane Zhang, X., Y. Han, Y. Sun, and Y. Li. 2019b. Innovative utilization of refractory
Science 5:223–66. iron ore via suspension magnetization roasting: A pilot-scale study. Powder
Uwadiale, G. O. 1992. Flotation of iron oxides and quartz-a review. Technology 352:16–24. doi:10.1016/j.powtec.2019.04.042.
Mineral Processing and Extractive Metallurgy Review 11 (3):129–61. Zhang, X., Y. Han, Y. Sun, Y. Lv, Y. Li, and Z. Tang. 2019a. An novel
doi:10.1080/08827509208914209. method for iron recovery from iron ore tailings with
Varicheva, A. V., A. A. Ugarova, N. T. Efendieva, S. I. Kretovb, A. A. Lavrinenkoc, pre-concentration followed by magnetization roasting and magnetic
A. A. Soludukhind, and P. V. Puzakovb. 2017. Innovative solutions in iron ore separation. Mineral Processing and Extractive Metallurgy Review 1–13.
production at Mikhailovsky Mining and Processing Plant. Journal of Mining doi:10.1080/08827508.2019.1604522.
Sciences 53 (5):925–37. Zhang, X., Y. Zhu, Y. Xie, Y. Shang, and G. Zheng. 2017. A novel macro-
Vidyadhar, A., K. H. Rao, and I. V. Chernyshova. 2003. Mechanisms molecular depressant for reverse flotation: Synthesis and depressing
of amine-feldspar interaction in the absence and presence of alco- mechanism in the separation of hematite and quartz. Separation and
hols studied by spectroscopic methods. Colloids and Surfaces A: Purification Technology 186:175–81. doi:10.1016/j.seppur.2017.05.051.
Physicochemical and Engineering Aspects 214 (1–3):127–42. Zhao, C., Z. Yahui, and C. Yongdan. 2012. Reverse flotation of quartz
doi:10.1016/S0927-7757(02)00361-8. from magnetite ore with modified sodium oleate. Mineral Processing
Vidyadhar, A., K. H. Rao, I. V. Chernyshova, and K. S. E. Forssberg. 2002. and Extractive Metallurgy Review 34 (5):320–30.
Mechanisms of amine–Quartz interaction in the absence and presence of Zhou, T. 2017. Research and application utilization status of the iron
alcohols studied by spectroscopic methods. Journal of Colloid and Interface ore flotation reagents. Modern Mining 6:98–102. (in Chinese).
Science 256 (1):59–72. doi:10.1006/jcis.2001.7895. Zhu, J. 1994. Iron ore processing technology of China. Beijing:
Vidyadhar, A., N. Kumari, and R. P. Bhagat. 2012. Adsorption mechan- Metallurgical Industry Press. (in Chinese).
ism of mixed collector systems on hematite flotation. Minerals Zhu, Y., B. Luo, C. Sun, J. Liu, H. Sun, Y. Li, and Y. Han. 2016. Density
Engineering 26:102–04. doi:10.1016/j.mineng.2011.11.005. functional theory study of α-Bromolauric acid adsorption on the α-quartz (1
Villar, J. W., and G. A. Dawe. 1975. The Tilden mine-a new processing 0 1) surface. Minerals Engineering 92:72–77. doi:10.1016/j.mineng.2016.
technique for iron ore. Mining Congress Journal 61 (10):40–48. 03.007.

View publication stats

You might also like