Sei sulla pagina 1di 26

Available online at www.sciencedirect.

com

Int. J. Miner. Process. 85 (2008) 59 84 www.elsevier.com/locate/ijminpro

Physical and thermal treatment of phosphate ores An overview


Abdel-Zaher M. Abouzeid
Cairo University, Faculty of Engineering, Department of Mining, Giza, Egypt Received 16 May 2007; received in revised form 6 September 2007; accepted 13 September 2007 Available online 19 September 2007

Abstract The annual consumption of phosphate rock approached 150 million tons. The marketable phosphate is usually 30% P2O5 or higher. The run-of-mine material is mostly of lower grade which needs processing or upgrading. The processing techniques of phosphate ores depend mostly on the type of associated gangue minerals present in the mined rock. In some cases, simple, inexpensive techniques are enough to produce the required grade. For example, crushing and screening is used to get rid of the coarse hard siliceous material, and attrition scrubbing and desliming is used to remove the clayey fine fraction. If silica is the main gangue material, single-stage or double-stage flotation is the conventional mineral processing technique used in this case. If the ore is igneous carbonatitic alkaline or ultra basic phosphate deposit, crushing, grinding, scrubbing, and flotation associated with other steps such as magnetic and/or gravity separation is proved to be successful in upgrading this type of ore. The sedimentary phosphate ores having carbonate-apatite as the main phosphate minerals and containing carbonates (calcite and/or dolomite) represent a challenge in the field of phosphate concentration due to similarity in the physico-chemical properties of surfaces of the ore constituents. Also, if considerable amount of organic matter constitutes the main gangue material, upgrading of the ore becomes difficult. New flotation systems (techniques and reagents) are being developed to treat these challenging phosphate ores. Furthermore, calcination is another solution for upgrading these difficult-to-treat types of ores. However, calcination is indicted with some controversial drawbacks. This overview discusses and summarizes the State-of-the-Art and the existing efforts to overcome these problems and to produce a high-grade phosphate product suitable for fertilizers and other phosphate compounds. 2007 Elsevier B.V. All rights reserved.
Keywords: Physical and mineralogical properties of phosphate minerals; Phosphate upgrading; Physical treatment of phosphate; Thermal treatment of phosphate; Flotation; Physical separation; Magnetic separation; Elecrostatic separation; Calcination; Defluorination

Contents 1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1. Types of phosphate rocks . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2. Mineralogical, chemical and textural characteristics of phosphate minerals . . 1.3. Techniques and measurements used to identify phosphate minerals . . . . . 1.4. The main parameters of phosphate characterization . . . . . . . . . . . . . Upgrading of phosphates according to ore type . . . . . . . . . . . . . . . . . . Techniques for upgrading phosphate ores . . . . . . . . . . . . . . . . . . . . . . Modified techniques and newly developed chemical reagents for phosphate flotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60 61 61 61 61 63 64 69

2. 3. 4.

E-mail address: abdel.abouzeid@gmail.com. 0301-7516/$ - see front matter 2007 Elsevier B.V. All rights reserved. doi:10.1016/j.minpro.2007.09.001

60

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

Thermal treatment of phosphate ores . . . . . . . . . . . . . . . . . . . . . . . 5.1. Energy consumption in thermal treatment of phosphate ores. . . . . . . . 5.2. Limitations and benefits of high temperature treatment of phosphate ores . 5.3. Types of thermal treatment units . . . . . . . . . . . . . . . . . . . . . . 5.4. Industrial calcination plants. . . . . . . . . . . . . . . . . . . . . . . . . 5.5. Effect of substituting carbonates on the characteristics of carbonate-apatite . 5.6. Characteristics of calcined phosphate product . . . . . . . . . . . . . . . 6. Effect of calcination on physical properties of the phosphate rock . . . . . . . . 6.1. Unit cell dimensions and crystal size . . . . . . . . . . . . . . . . . . . . 6.2. Surface area, porosity, and material density . . . . . . . . . . . . . . . . 6.2.1. Surface area . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2.2. Porosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2.3. Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.3. Effect of calcination on chemical reactivity of the calcined product . . . . 7. Summary. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.

. . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . .

72 73 73 74 74 74 76 77 77 78 78 79 79 80 81 82 82

1. Introduction Phosphates, in the form of fertilizers, are essential in the agricultural sector. They are also very important constituents in animal feed stocks and in food and other chemical industries. About 95% of the world phosphate rock production is consumed in fertilizer industry. Most of the balance is processed, in electric furnaces, into elemental phosphorus which is the main raw material for manufacturing various phosphate compounds (Emich, 1984; Jasinski, 2007). Economic recovery of phosphates is limited to naturally concentrated phosphate mineral deposits. Occasionally natural concentrations are enough to be used as mined; generally, however, the ore is low

grade and must be concentrated for economic utilization. Physical as well as thermal treatment techniques are used for concentrating the run-of-mine phosphates. More than 40 countries around the world produce phosphate rock. In 2005, the world phosphate production reached 148 Mt, out of which 143 Mt were produced by sixteen major phosphate producing countries (countries whose production exceeds 1 million tons per year) (Jasinski, 2005, 2007). These countries, arranged in descending order of their production capacities, are: USA, Morocco, China, Russia, Tunisia, Jordon, Brazil, Israel, Syria, South Africa, Egypt, Australia, Senegal, India, Togo, and Canada. Only 5 Mt are produced by all other countries. Fig. 1 shows the growth curve of the world annual phosphate production since the

Fig. 1. World phosphate production rate since 1850 (Jasinski, 2007; Abouzeid et al., 1996).

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

61

year 1850, when the first world phosphate production was recorded (Jasinski, 2007; Abouzeid et al., 1996). 1.1. Types of phosphate rocks There are five major types of phosphate resources in the world (Straaten, 2002, 2007): Marine phosphate deposits. Igneous phosphate deposits. Metamorphic deposits. Biogenic deposits. Phosphate deposits as a result of weathering.

The world phosphate resources are distributed, according to their type, approximately as follows: 75% from sedimentary marine deposits, 1520% from igneous, metamorphic and weathered deposits, and 2 3% from biogenic sources (bird and bat guano accumulations). Fig. 2 shows the location of the major phosphate rock producers as well as some of the known phosphate deposits that were not yet in production (Emich, 1984). 1.2. Mineralogical, chemical and textural characteristics of phosphate minerals Phosphate minerals occurring in the primary environment include (Straaten, 2002; Abu-Eishah et al., 1991): Flour-apatite (Ca10(PO4)6F2), found mainly in igneous and metamorphic environments, for example, in carbonatites, and mica-pyroxenites. Hydroxy-apatite (Ca10(PO4)6(OH)2), found mainly in igneous and metamorphic environments but also in biogenic deposits, e.g. in bone deposits. Carbonate-hydroxy-apatite (Ca10(PO4,CO3)6(OH)2), found mainly on islands and in caves, as part of bird and bat excrements, guano. Francolite (Ca10 x yNaxMgy(PO4)6 z(CO3)zF0.4zF2). This complex carbonate-substituted apatite is found only in marine environments, and, to a much smaller extent, in weathered deposits, for instance over carbonatites. Dahllite 3Ca3 (PO4)2CaCO3. This phosphate structure is found in the marine sediments. Collophane 3Ca3 (PO4)2nCa(CO3,F2,O)xH2O. This type of phosphate minerals is typical of the marine phosphate sediments. The ability of anions to substitute for phosphate in apatite is not restricted to F, CO3 2 and OH, which are

found in the differing variants of apatite. Silicate, vanadate, aluminate, titanate and arsenate may also occur in place of part of the phosphate. Likewise, calcium may be replaced in part by uranium, lanthanide rare earths or certain other trace metals such as cadmium, zinc, magnesium, strontium and barium. In some cases chlorine occurs in place of part of the fluorine. The extent to which substitution takes place and what the substitutes are, depend on the prevailing conditions when the apatite was formed and on the influence of subsequent events, such as weathering. In sedimentary deposits of marine origin, the phosphate material occurs in admixtures with detritus materials such as quartz, mica and clay, often with limestone and occasionally with dolomite. Igneous apatites may contain other impurities, not commonly found in sedimentary deposits, originating from other constituents of the magma from which the apatite crystallized (USGS, 2003). 1.3. Techniques and measurements used to identify phosphate minerals The most important techniques used for identifying phosphate minerals are (Straaten, 2002): X-ray diffraction (XRD) for mineralogical analysis. X-ray fluorescent (XRF) for elemental analysis. Scanning Electron Microscope (SEM) for grain interaction and spot analysis. Infra Red methods (IR) for detection of surface-layer structure. Micro-probe analysis for elemental distribution in the matrix. Microscopic identification for mineralogy and grain interrelation. Analytical chemical methods. Measurement of refractive indices. Porosity and pore size distribution measurements. Surface area measurements. Thermal Analysis (DTA), (TGA). Unit-cell dimensions (a-value), and crystal size. 1.4. The main parameters of phosphate characterization There are several parameters that are used in characterizing phosphate rocks. Among these parameters are (Straaten, 2002). Unite-cell a-dimension (a-value), a crystallographic expression of the apatite composition. The a-cell dimension is expressed in Angstrom (1 = 10 10 m). Refractive index.

62

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

Fig. 2. Location of the major phosphate rock-producing areas of the world and some of the known deposits not now in operation (Emich, 1984).

Solubility data based on chemical extraction methods, e.g. neutral ammonium citrate solubility (NAC). PO4/CO3 ratio as a measure of carbonate substitution in phosphate minerals. Surface area (expressed in m2 g 1), and pore size distribution indicating potential reactivity. The calcium phosphate content of phosphate rock is expressed in different world areas by one of the following terms (Emich, 1984): BPL TPL P2O5 P (bone phosphate of lime) (triphosphate of lime) (phosphorus pentoxide) (phosphorus not commonly used)

An illustration of relationship between the above terms is: 80% BPL = 80% TPL = 36.6% P2O5 = 16% P. High-grade phosphate rock is available from several sources, such as Togo, Senegal and Morocco. However, premium rocks from these sources are not available at costs comparable to the past few decades. It has been widely suggested that, in general, there have been a continuous decrease in world phosphate rock quality as

reserves of high-quality rock are being depleted. As the high-grade phosphate deposits are becoming depleted, the quality of phosphate rock that is utilized on a worldwide basis is decreasing. High P2O5 content equates to low impurity content, high yields per ton of material shipped, handled and processed, increased reaction efficiencies, fewer processing problems and less waste. An industry assessment suggests that the phosphate content of pre-beneficiated ore is already decreasing by around 1% per decade (USGS, 2003). Estimates of world phosphate reserves and availability of exploitable deposits vary greatly and, assessments of how long it will take until these reserves are exhausted vary also considerably. Furthermore, it is commonly recognized that the high-quality reserves are being depleted expeditiously and that the prevailing management of phosphate, a finite non-renewable source, is not fully in accord with the principles of sustainability. For example, if the world population increases by 50% over the next 50 years, this would point towards an increase in global food needs by at least a proportionate figure, assuming roughly constant per capita cereal consumption. As demand for food increases, this may result in bringing into agricultural use more land, but certainly will bring a

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

63

requirement for increased yields, thus increasing fertilizer demand. Hence, agricultural phosphate use may increase faster than world population (Haarr, 2005). Depletion of current economically exploitable reserves are estimated at somewhere from 60 to 130 years. Using the median reserves estimates and under reasonable predictions, it appears that phosphate reserves would last for at least 100+ years. Increasing demand and increasing prices will make more reserves economically exploitable (Steen, 1998; Evans and Johnston, 2004). Based on the present rate of production, about 150 Mt/a, and the available extractable reserves, 18 billion ton, the life time of these known reserves is about 120 years (Jasinski, 2007). 2. Upgrading of phosphates according to ore type Because of the increasing demand on phosphate rock for fertilizers, it is becoming more and more common to mine and process low grades of phosphate deposits. The majority of these low grade deposits are of sedimentary phosphates and the rest are of different origins. These deposits occur in the form of ore bodies in one thick bed or several successive beds intercalated with non-phosphatic materials. Phosphate ores show a wide diversity in the composition of their gangue materials but generally fall into one of the following categories, based on the major associated gangue materials (Abouzeid et al., 1980): (1) Siliceous ores: These contain quartz, chalcedony or different forms of silica. Such ores could be upgraded economically by such techniques as flotation or gravity separation methods. (2) Clayey ores: These mainly contain clays and hydrous iron and aluminum silicates or oxides as gangue materials. These impurities could be removed by simple beneficiation techniques such as scrubbing and washing. In some cases, dispersing agents would be necessary. (3) Calcareous ores of sedimentary origin: These contain calcite and/or dolomite as the major impurities with small amounts of silica. It is usually difficult to remove the carbonate minerals efficiently from such ores by conventional techniques as flotation (Cathcart, 1967; Service and Popoff, 1967; Elgillani et al., 1984), or by physical separation methods. This is because the physical properties of carbonates and phosphates are very similar. Separation by physical means even becomes impossible when the carbonate minerals are finely disseminated into the phosphate particles. Experimental results carried out by Wilson and

Ellis (1984) showed that the flotation techniques applied on a calcareous ore may be effective only if the phosphate can be highly liberated from the gangue materials in a relatively coarse size. In the case of the unaltered ores, complete liberation is achieved only with extremely fine grinding. Consequently, partially liberated flotation feed must be used to avoid loss of selectivity resulting from fine grinding. Under such conditions high concentration ratios cannot be reached. For upgrading such type of ores, crushing and screening may be used if there are differential friability differences between phosphates and carbonates. However, recovery in this case may be unreasonably low. An alternative technique for upgrading these ores is calcination. Calcination is the process of heating the ore to a high temperature ranging from 800 to 1000 C to decompose the CaCO3 and MgCO3 to CaO, MgO and gaseous CO2. The CaO and MgO formed are then removed as hydroxides by quenching the calcined product in water and washing. The most common chemical reagent used to enhance the removal of calcium and magnesium hydroxides is the ammonium chloride, NH4Cl (Orphy et al., 1969; Ahmed, 1990; Gu, 2002). Chemical dissolution of carbonate minerals (calcite and dolomite) from calcareous phosphate ores, without calcination, using organic acids also proved capable of upgrading the calcareous phosphate ores on the laboratory scale (Gu, 2002; Abu-Eishah et al., 1991; Sengul et al., 2006; Ahmed, 1999; Zafar et al., 2006). (4) Phosphate ores associated with organic matter, which sometimes called black phosphates (Blazy and Bouhaouss, 2005): Ores of this type are generally upgraded by heating the ore up to about 800 C. This type of calcination burns the organic material and the residual organic carbon without significantly affecting the superior qualities of the sedimentary phosphates such as solubility and reactivity. Further more, as a result of the low calcination temperature; the reduction of calcium sulfate, present in the ore, to corrosive calcium sulfide by the organic matter is minimized. During burning the organic matter, the following two conditions must be satisfied: organic carbon must be decreased to less than 0.3% to minimize gassing in the wet phosphoric acid processing, and apatite CO2 must be maintained at a level close to 2% to allow good reactivity to the calcined product (Blazy and Bouhaouss, 2005; Henin and Lectard, 1983; Henin and Pinoncely, 1986).

64

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

(5) Phosphate ores containing more than one type of the gangue minerals: Many of the sedimentary phosphate deposits contain mixtures of undesired constituents. These ores require series of upgrading operations in their flow sheets depending on the type of gangue minerals existing in each ore. This may include, after size reduction, combination of attrition scrubbing, desliming, flotation, gravity separation, and/or calcination. Each flow sheet is to be designed after thorough characterization of and testing on a representative sample of the ore under exploitation. (6) Igneous and metamorphic phosphate ores: The main gangue materials in these ores are sulfides, magnetite, carbonates (calcite, dolomite, siderite, and ankerite), nepheline syenite, pyroxenite, foskorite, etc. Processing of these ores may include, after crushing and grinding, washing, desliming, magnetic separation, and flotation depending on the types of gangue minerals present. However, flotation is a common step in all of them (Emich, 1984). About 20% of the world's phosphate rock now comes from the mining and beneficiation of igneous deposits, but this percentage will drop as production from sedimentary deposits increases. Igneous apatite in relatively small quantities has been recovered from veins and from magnetite ore bodies where it occurs in minor amounts. The important igneous apatite deposits of commercial interest are found in certain intrusive complexes of alkalic rocks. Commonly near the center, there are veins and larger bodies rich in carbonates (calcite, dolomite, siderite, and ankerite) emplaced either in a molten condition or by metasomatic processes. The calcareous deposits of the igneous phosphate rocks are called carbonatites. The assay of phosphate rocks of the igneous deposits is generally in the range of 7 15% P2O5. World production of apatite (igneous) concentrates in 1980 was about 25 Mt. Table 1 lists most of the operating apatite mines in the world in 1985. Brazil is actively developing new apatite mines; in 1985 its production of phosphate rock, largely igneous apatite, was estimated to
Table 1 World apatite (igneous) mines in 1985 (Emich, 1984) Country USSR (Kola) South Africa (Phalaborwa) Brazil Zimbabwe (Dorowa) No. mines 6 2 5 1 No. beneficiation plants 2 1 3 1

be 5.8 Mt, compared to 1.7 Mt in 1980. The Russians developed new operations in the Kola area and have new mines operating on apatite in the Lake Baikal region in Siberia. Some of the Russians apatite mines in the Kola Peninsula are underground, but the majority of its mines, as well as all other apatite mines in the world, are surface operations. Foskor, in South Africa, developed additional large mine-beneficiation operations on the Phalaborwa carbonatite complex. Carbonatites have been found in Australia in 1968, New Zealand in 1970, and early in 1980s in Paraguay (Emich, 1984). 3. Techniques for upgrading phosphate ores There are several beneficiation techniques for upgrading phosphate ores. The choice of one or more of these techniques depends on the type of ore as well as the associated gangue minerals. Among these techniques are: 1) Size reduction and screening: This technique utilizes the differences in differential friability between phosphate minerals and associated gangue and cementing matrix (carbonates, silica, and silicates). In most cases, phosphate minerals are friable while gangue minerals are hard. By crushing and screening, the fine fraction will be rich in phosphates, but high percentage of P2O5 is lost in the coarse fraction. Air elutriation may be practiced to get rid of the very fine fraction that is rich in clays. An example of using this technique is the beneficiation plant of the Nile Valley phosphate rock at East Siba'eya area, Egypt. There the main beneficiation facilities are crushing and screening to recover the phosphate product in the fine fraction, but the coarse fraction, the reject, may contain up to 22% P2O5 (Negm and Abouzeid, 2004). 2) Attrition scrubbing and classification: Examples of the operations using this technique are listed below: a) This technique is used when the main gangue minerals are clays. Clays are characterized by their fine size, loosely bound grains with the phosphorite pelets. Attrition in water liberates and disperses the clay particles in water and they are removed by desliming

Concentrate, 1000 t/a Grade, % P205 20 3.5 5.8 1 17 7 13.515 7

Type deposit Nepheline syenite (unweathered) Pyroxenite and foskorite (unweathered) 1 carbonatite (unweathered) 4 carbonatite weathered residuals. Alkalic complex.

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

65

and/or classification. Such technique is successfully used to upgrade some of the Red Sea Coast phosphate rocks in Egypt (Khalifa et al.,1972). b) It is also used when the associated gangue minerals are mainly coarse silica or cherty chips. In this case, phosphorites are collected from under the screens in a wet process. This may also be practiced to separate a coarse phosphate-rich fraction ahead of a flotation section such as in Florida plants (Emich, 1984; EPA, 1993). 3) Electrostatic separation: In combination with attrition, desliming, and gravity separation, the electrostatic separation technique was successful in upgrading phosphate ores by removing silica and/or carbonates, mostly on laboratory scales. However, the low capacity of the electrostatic separators limits their use for large scale production. This technique is used to concentrate phosphate ores of different types. The following are examples of the ores responded to the electrostatic separation: a) A calcareous ore (from North Africa) was upgraded using electrostatic separation to produce a concentrate assaying: 32.9% P2O5 at 80% P2O5 recovery from a feed assaying 24.1% P2O5 (Ciccu et al., 1972). b) Abu-Tartour ore, Eastern Desert, Egypt (containing quartz, ankerite, gypsum and pyrite), produced a concentrate assaying 33% P2O5 at a recovery of 71.5% from a feed assaying 27.5% P2O5 (Hammoud et al., 1977). c) A siliceous ore (from Qusseir, Red Sea Coast, Egypt) produced a concentrate assaying 33.8% P2O5 at 70.3% recovery from a feed assaying 16.5% P2O5 (Abdel Moneim, 1971). d) A siliceous ore (from Hamadat mine, Red Sea Coast, Egypt) produced a concentrate assaying 30% P2O5 at a recovery of 76.3% from a feed assaying 18.2% P2O5 (Abouzeid et al., 1996). e) One of the few plants that used electrostatic separators on an industrial scale was the Pierce plant in Florida, USA. The plant used to upgrade phosphate flotation-concentrate using electrostatic separators. The feed assayed 32.9% P2O5 and 11.3% insoluble, and the final concentrate assay was 35.3% P2O5 and 5.6% insoluble (Taggart, 1964). 4) Magnetic separation: When one of the major gangue constituents is magnetic, magnetic separators are used as one of the steps in the flow sheet to remove the magnetic constituents. This is mostly used in the beneficiation of igneous phosphate rocks. However, it has also been used for upgrading some sedimentary phosphate ores. The following are examples of plants using magnetic separation in their flow sheets:

a) The complex igneous phosphate deposit at phalaborwa (South Africa) is upgraded in three stages (Lovell, 1976): 1. copper flotation, 2. magnetic separation, 3. the non-magnetic fraction is subjected to flotation to recover apatite. The Run-Of-Mine, ROM, to this plant assayed 7% P2O5, and the phosphate concentrate assay was more than 36% P2O5 at a recovery of 7580% P2O5. b) In Brazil's long established apatite (igneous) beneficiation plants, magnetic separation followed by desliming and flotation are included in the phosphate concentration plants. Ores assaying 7 15% P2O5 are upgraded to produce concentrates assaying 36 to 38% P2O5 (Silvia and Andery, 1972; Breathitt and Finch, 1977). c) The dolomitic-pyritic sedimentary phosphate ore of Abu-Tartour phosphate deposit in Egypt is subjected to high intensity magnetic separation, after desliming, to separate the feebly magnetic ankerite. The feed assays about 2226% P2O5 and the phosphate concentrate assays 2930% P2O5 at a recovery of 70% (MIMW, 2000). d) The Sukulu weathered carbonatite deposit, Uganda, which has an average grade of 12.8% P2O5 produced a phosphate concentrate assaying 4042% P2O5. The mineralogical constituents of this ore were approximately 32% apatite, 57% magnetite and goethite, and 0.25% pyrochlore. The Sukulu apatite ore was concentrated through grinding and magnetic separation followed by flotation (Straaten, 2002). 5) Chemical dissolution of carbonates: Calcium and magnesium carbonates are readily dissolvable in both mineral acids (strong acids) and organic acids (weak acids). In the case of calcareous phosphate ores, although mineral acids dissolve carbonates at high reaction rates, they also attack the phosphorus bearing minerals and cause losses in the P2O5 content of the ore if the intention is only to upgrade the ore not to dissolve the phosphates. To avoid this problem, organic acids have been studied as carbonate leaching agents although their reaction rates are slow. These organic acids may be expensive and will certainly add to the production cost. On the other hand, they are selective to leaching carbonates, their capital cost is low, they do not cause environmental hazards, and they can be recycled (Ahmed, 1999; Gu, 2002). Any way, the overall economic analysis of the operation decides whether the leaching process could be used or not (Gu, 2002).

66

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

The organic acids most commonly used in carbonate leaching are: acetic acid, citric acid, and formic acid. For some specific advantages (may be cost, availability, etc.), acetic acid attracts the attention of many researchers (Gu, 2002). The suggested reaction between acetic acid and carbonates is as follows (Gu, 2002; Sengul et al., 2006): CaCO3 2CH3 COOH CaCH3 COO2 CO2 H2 O The acetic acid may be recovered by reversing the above reaction at high CO2 pressure in a separate reactor, or by using sulfuric acid to precipitate calcium sulfate and to free acetic acid (Sengul et al., 2006; Gu, 2002). The most important parameters investigated in the area of carbonate leaching from phosphate ores by organic leaching agents are: particle size distribution of the feed, type and nature of the ore, solid/liquid ratio, reaction time, leaching reagent, reagent concentration, stirring speed (this parameter is related in some way to leaching time, in addition to the rate of attrition which is a function of time and speed), and temperature (Sengul et al., 2006; Gu, 2002). It was found that room temperature is the optimum temperature for this type of reaction because higher temperatures tend to decrease the solubility of the acetate which makes separation difficult (Abu-Eishah et al., 1991), and increases the dissolved P2O5 (Sengul et al., 2006). The dissolution time in most of the studied cases ranged from 20 to 40 min. The following are some results of laboratory experimental work carried out for upgrading calcareous phosphate ores from different locations using organic acids as leaching agents: a) Abu-Eishah et al. (1991) used different types of Jordanian ores (soft, 2 mm, 29% P2O5, 10% LOI; hard, + 2 mm, 20.1% P2O5, 20% LOI; and combined mixtures of both at different ratios). The feed to the reactor was 1.981 mm. They obtained concentrates assaying 3234.3% P2O5 and 3.64.2% CO2 depending on the type of feed to the reactor. The acetic acid concentration was about 6% at liquid/solid ratio of 5:1. b) A phosphate sample from Um Hammad area, Eastern Desert, Red Sea Coast, Egypt, assaying 25.2% P2O5 and 12.3% CO2 was leached by an 8% acetic acid solution at liquid/solid ratio of 6:1 (Ahmed, 1999). The feed was 1 mm size. The obtained concentrate assayed 32.2% P2O5. c) Using 0.5 M acetic acid solution, Sengul et al. (2006) leached a calcareous phosphate sample from MardinMazidagi Kasrik, Turkey, at room temperature. The

feed was 250355 m size and assayed 1013% P2O5 and 3032.6% CO2. The obtained concentrate assayed 2930% P2O5. d) Zafar et al. (2006) leached a phosphate sample from Pakistan using formic acid. They reported that under optimum operating conditions the P2O5 was raised by 3035% with corresponding reduction in CO2 by up to 6772% at a recovery of 70% by weight. They confirmed that increasing temperature led to decreasing the solubility of the produced calcium formate. e) Treating a different phosphate sample, Zafar et al. (1995) used a dilute formic acid solution to leach the carbonates in the concerned ore sample. The P2O5% was increased by about 29% with a reduction of about 69% in the carbonate content of the feed sample. In another publication, Zafar (1993) showed that P2O5% was increased by 23% at a corresponding reduction in CO2 by about 62%. The above results show that organic acids can effectively reduce carbonate content in phosphate ores. This suggests that if economy favors upgrading of a phosphate ore by organic acids leaching, this technique should be seriously considered. It is less environmental hazardous because organic acids are degradable. 6) Flotation: More than 60% of the marketable phosphate in the world is produced by flotation (USGS, 2007). Siliceous ores: The common technique for processing siliceous phosphate ores is the anioniccationic flotation technique, the Crago technique. If there is any carbonate minerals, it will report to the phosphate concentrate (Hanna, 1964; AZCC, 1981; Hanna and Anazia, 1990). Calcareous ores: phosphate minerals and carbonate minerals (Calcite and dolomite) exhibit very similar surface chemical properties. Sedimentary phosphates are more sensitive to the similarity in surface chemical properties of these minerals. This makes their separation by flotation a challenging problem Elgillani, 1978/1979; Elgillani et al., 1984; El-gillani, 1992; El-gillani and Abouzeid, 1993). Much research has been carried out to try to reduce the calcium carbonate content of phosphate rocks by flotation and/or calcination (Zafar et al., 1995). Hignett et al. (1977) claimed that flotation seems to work best on ores containing well-crystallized carbonates. When the ore contains soft or chalky carbonates the flotation results were less satisfactory. This was the case of East Mediterranean and North Africa ores where the carbonate crystals are intergrown so that the phosphate is not liberated by comminution until the rock is too fine for flotation (Mew, 1980; Baudet, 1988; Crago, 1940).

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

67

However, as discussed previously, it is obvious that the problem of separating carbonates from phosphates in the sedimentary rocks is seriously complicated, and a brief survey on the latest efforts for solving this problem is due. The currently used commercial process of double flotation, the Crago process, is not adequate for reducing the dolomitic impurity level to less than 1.0% by weight MgO in the concentrate, as stipulated by the phosphate industry. During the past few decades, studies conducted by investigators in the Florida phosphate industry (Lawver et al., 1978, 1982; Snow, 1979; Dufour et al., 1980) and the Tennessee valley Authority (Lehr and Hsieh, 1981) have resulted in the development of a number of processes. However, all these studies originated on a bench scale and the fundamentals involved in the separation scheme have not been well established. This leads to serious limitations for optimizing these processes and also for suggesting the specific conditions to be used for beneficiating ores of different origins and compositions. Recently, Moudgil and Chanchani (1985, 1986) and Ince (1987) have conducted some fundamental studies which resulted in the development of two processes for the removal of dolomite from apatite. Extension of these fundamental studies to beneficiate the natural ores on a bench scale was also reported (Moudgil et al., 1990). Also, Zhong et al. (1991) have attempted a systematic optimization of some variables using a three stage flotation scheme with various reagents. To help understand the collector adsorption mechanisms in the phosphate-carbonate systems, more extensive investigations were conducted to clarify the effect of the addition of organic and inorganic salts on the surface charge of the two species, apatites and carbonates (Doss, 1976; Bell et al., 1973; Mishra et al., 1980; Mishra, 1978; Smany et al., 1975; Somasandaran and Agar, 1972). These studies established firmly that phosphate ions function as potentialdetermining ions for apatites, while Ca++ ions act as either potential-determining ions (Eigeles, 1958; Saleeb and DeBruyn, 1972) or specifically adsorbed ions (Somasandaran, 1968; Aplan and Fuerstenau, 1962; Hanna and Somasandaran, 1976) for apatites. In the case of calcite, both Ca++ and carbonate ions have been shown to be potential-determining ions for the mineral (Mishra, 1978; Smany et al., 1975; Somasandaran and Agar, 1976). Flotation behavior of apatites and carbonates was conducted under various experimental conditions (Saleeb and DeBruyn, 1972; Somasandaran, 1968; Aplan and Fuerstenau, 1962; Hanna and Somasandaran, 1976; Johnston and Leja, 1985). Many et al. (1975) provided a comprehensive review of electrophoretic studies of phosphates and associated calcareous gangue minerals, together with flotation results

of selective separation of phosphate oolites and calcite. In their investigation, it was shown that H+, OH, Ca++, and HPO4 2 function as potential-determining ions for apatites. They also demonstrated that adsorption of sodium dodecylsulfonate and dodecylamine chloride on phosphate surface is due to electrostatic attraction while adsorption of oleate at pH greater than 6 is due to a chemical binding energy between oleate ions and surface sites of the phosphate particles. Between about pH 5.5 and 6.5, the phosphated-surface particles were sharply depressed, whereas their flotation recovery increased at pH 4. Depression of such particles between pH 5.5 and 6.5 was explained to be due to the lack of adsorption of the collector, whatever the concentration may be. The flotation at pH 4 was suggested to be due to adsorption of neutral molecules in the vicinity of the isoelectric point, IEP. Mitzmager and Co-workers (1966) dealt with the effect of soluble alkali phosphate salts on oleate flotation of calcite from phosphate slimes. They tested the selective depressing action of soluble alkali phosphate salts on phosphate minerals. The authors concluded that the mechanisms of the depression of phosphate and activation of calcite by soluble alkali phosphate are due, at least, to two parallel effects: a) The adsorption of molecular dicalcium phosphate (CaHPO4) on the phosphate particles probably as a replacement of dissolved gypsum, and b) The mild attack on calcite particles accompanied by the evolution of tiny CO2 bubbles on the surface. Johnston and Leja (1985) continued the effort to understand the complex nature of the phosphatecarbonate systems. Flotation behavior of two apatite samples and a dolomite sample were investigated in presence of oleic acid and soluble phosphate solutions. They concluded that: (1) Flotation of apatite is prevented by the depressing action of soluble phosphate ions in acid solutions, whereas dolomite floats readily in their presence. (2) If gypsum is present in the flotation feed, excessive quantities of phosphate ions are needed for the depression of apatite because of a fast metathesis reaction between gypsum and orthophosphate ions. This reaction can be completely suppressed by the addition of suitable quantities of sulfate ions, which is exactly opposite to the findings of Mitzmager and co-workers (1966). (3) Following the findings reported by Bertolucci (1968), it was suggested that the depressing action of phosphate ions on apatite is due to the formation of stronger hydrogen-bonded phosphate-water

68

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

layer around apatite particles than around dolomite particles. Evolution of CO2 from dolomite or calcite, in acidic solutions, disturbs the hydrogen bonding around these particles. This allows an immediate adsorption of collector and results in dolomite or calcite flotation. Moudgil and Chanchani (1986) used two-stage conditioning process for the flotation of dolomite from apatite. The mixture of minerals (dolomite + apatite) was first conditioned at pH 10 in presence of a fatty acid collector. The system was then reconditioned below pH 4.5 where dolomite was floated. Soto and Iwasaki (1985) showed that in a mixture of francolite and dolomite, the primary amines were uniquely selective towards the phosphate mineral. In a phosphate ore, primary amines were adsorbed mainly on the phosphate indicating that stronger interaction existed between the amine and francolite than between the amine and dolomite. The US Bureau of Mines developed a process to separate carbonate (calcite) from phosphate minerals occurring in the unaltered phosphate ores of Phosphate Formation (Rule et al., 1974). The process uses fluosilicic acid to depress the phosphate while floating the calcite with fatty acid emulsion. Another US Bureau of Mines report of investigation dealt with the beneficiation of high-magnesium phosphate from Southern Florida (Llewellyn et al., 1982). Several depressants such as starches of different sources, fluosilicic acid, sodium borate, lactic acid, sodium biphosphate, phosphoric acid, and acetic acid failed to depress phosphate minerals. A technique of sizing, grinding, desliming, scrubbing, redesliming, and flotation was then devised to treat three southern Florida high-MgO phosphate ores. Moudgil, et al. (1986) described a flotation process based on the depression of apatite with sodium chloride in acid medium and flotation of dolomite with fatty acids. When this process was applied to a natural phosphate ore, it was possible to float dolomite from apatite at pH = 4. Hanna and Anazia (1990) described a flotation process developed by Alabama Mineral Resources Institute (AMRI) for selective fatty acid flotation of carbonates from sedimentary apatites in the pH range of 46 without conditioning of the pulp prior to flotation. The success of the process is based on the fact that fatty acids adsorb more rapidly onto carbonate surfaces than onto phosphate surfaces. Carbonates are floated in slightly acidic circuit immediately upon the addition of the collector and frother, then the pulp is conditioned briefly, and the phosphates are floated from the siliceous constituents with no further addition of collector in most instances. Important as they are, these studies have not provided an understanding of the mutual effects of phosphate and

carbonate minerals on each other during conditioning and flotation. This shows the need for a thorough thermodynamic analysis of the phosphate-carbonate-water systems in presence and absence of flotation reagents. Simultaneously, in 1978/1979, two studies were published by Johnston and Leja (1978) and Elgillani (1978/ 1979) on the analytical aspects of the calcite/apatite/water system. These analyses showed that the control of dissolved Ca++ in the system seems to be the key to achieve selective flotation of carbonates, especially calcite, from apatite. Based on thermodynamic analysis of the apatite / calcite / water system that was carried out in these studies, it was concluded that: (1) Depression of apatite is probably related to its instability in acid media. (2) Concentrations of aqueous calcium phosphate + complexes such as CaHPO4 and CaH2PO4 increases with the decrease in pH of the system. (3) The phase boundary line between hydroxyl-apatite and CaHPO4(S) occurring at pH 4.6 is shifted towards more acidic pH values when dissolved Ca++ is added to the system and towards less acidic pH values when soluble phosphate salts are added to the system. This may explain why it is relatively easy to depress apatites in acid media in absence of carbonates, while the same thing cannot be achieved when natural ores are treated. These theoretical analyses were extended a step further (El-gillani, 1992) in order to assess the most important conditions for separation of carbonates from phosphates. The final conclusion of this study was that: Concentration of calciumphosphate complexes, in the solution, in equilibrium with all apatites increases with the decrease of pH and with the increase of added soluble phosphate, oxalate, and sulfate salts; but sharply decreases by adding or dissolving calcium salts. To sum up this theoretical discussion, two conditions seem essential to be fulfilled in order to achieve selective flotation of carbonates from apatites in acid media (pH 5.56.0). These are: minimizing Ca++ and/or increasing HPO4 2 in the carbonate/ apatite/water system (El-gillani and Abouzeid, 1994). The following are examples for separating carbonates from phosphates, on the laboratory scale, by flotation: a) The USBM treated a sedimentary ore (carbonateapatite in addition to pyrite, Ca and Mg carbonates and silica). Three-stage flotation operation was adopted starting with flotation of pyrite using

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

69

xanthates. They added H2SiF6 as depressant for phosphate minerals, and used anionic collector to float the carbonates followed by cationic flotation to remove silica. The phosphate concentrate assayed 29% P2O5 at 69% recovery from a feed assaying 19.9% P2O5 (see Fig. 3) (Rule et al., 1978). b) A thermodynamic analysis of the apatite-calcite system was carried out at Cairo University (El-gillani and Abouzeid, 1993). It showed that at pH= 5.8, it is possible to float carbonates from calcareous sedimentary phosphate ores. A flotation scheme was designed such that the pH value was kept constant around 5.8 during conditioning and flotation of a deslimed sample. Flotation continued until barren froth was observed. Both the carbonate float and phosphate concentrate were filtered, washed, dried, weighed, and analyzed. Using this scheme, the following results were obtained for samples from two mines from the Red Sea Coast phosphate rock, Egypt (see Tables 2 and 3, feed size range 125 + 25 m, pH range 5.75.9).

4. Modified techniques and newly developed chemical reagents for phosphate flotation The conventional phosphate flotation process is the Crago flotation technique. This technique involves a cationicanionic sequence of operations. In the anionic step, the phosphate ore is conditioned using fatty acid/fuel oil as collectors and then the phosphate minerals are floated in alkaline media leaving the majority of silica and silicate minerals as non-floats, tailings. The phosphate concentrates are scrubbed at pH 4 to remove the anionic collector, and then subjected to cationic flotation in an acidic solution to float the silica entrapped within the phosphate concentrate during the anionic step. However, 3040% by weight of the silica present in the feed are floated twice. An improvement of the second stage of the Crago process, cationic stage, was suggested by Schreider, (1981). The cationic reagent was added in two dosages in starvation amounts, with a low retention time to float most of the silica. Additional cationic reagent is added with a retention time sufficient to produce

Fig. 3. Beneficiation of partially altered sedimentary phosphate rock containing carbonates, silica, and pyrite as the main gangue material (Rule et al., 1978).

70

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

Table 2 El-Hegab El-Shemali Phosphate Sample (the feed assayed 23.0% P2O5) (El-gillani and Abouzeid, 1993) Product Carbonate float Phosphate conc. Carbonate float Phosphate conc. Carbonate float Phosphate conc. Carbonate float Phosphate conc. Carbonate float Phosphate conc. Carbonate float Phosphate conc. Carbonate float Phosphate conc. Carbonate float Phosphate conc. P2O5, % 16.9 28.7 16.92 30.85 14.57 31.51 15.1 32.99 16.2 30.1 15.0 33.4 18.0 30.66 17.39 31.26 Insoluble, % 2.71 3.12 2.61 2.87 2.53 3.10 2.96 3.22 2.84 2.95 2.72 2.93 2.94 1.48 2.22 1.1 LOI, % 27.34 12.42 27.39 10.63 28.96 10.51 28.6 9.42 27.52 10.93 28.87 9.2 22.62 8.71 23.66 8.5 P2O5Rec., % 65.1 62.37 67.75 67.18 64.00 64.7 61.9 64.9 Modifying reagents a 2.4 kg/t H2 SO4 a 2.8 kg/t H2 C2 O4 a 2.1 kg/t H3 PO4 a 6.4 kg/t KH2 PO4 b 1.6 kg/t KH2 PO4 b 1.6 kg/t H2 SO4 1.6 kg/t KH2 PO4 b 2.4 kg/t H2 C2 O4 5.1 kg/t H3 PO4 c 1.7 kg/t H2 SO4 5.7 kg/t H3 PO4 c 1.9 kg/t H2 C2 O4

Phosphate depression and pH controller, collector was 0.6 kg/t in all experiments. a Reagent was added during conditioning and flotation. b KH2 PO4 was added before conditioning and pH was adjusted by H2 SO4 and/or H2 C2 O4. c Reagents were added as mixed solutions during conditioning and flotation.

a clean phosphate product. Approximately two thirds of the phosphate lost in the conventional flotation technique is recovered by this process. To avoid the drawbacks of the Crago process, the Florida Institute of Phosphate Research, FIBR, developed a cationicanionic flotation scheme, which is the reverse of the Crago process. In this process, fine sands are first floated with a minimum dosage of amine condensate added stepwise. The phosphate concentrate from this step is then floated with a blend of surfactant/fatty acid/fuel oil collector. This process simplifies the current Crago flow sheet by eliminating the acid scrubbing circuit and reducing the number of conditioners and the amounts of

the reagents down to one-third to one-half of the amounts used in the Crago process at the same efficiency (grade and recovery) (Zhang, et al., 1997). There is also the all-cationic flotation technique. This process consists of using a relatively cheap amine condensate with sodium tripolyphosphate depressant to float fine sands from the 14 150 mesh feed. The concentrate is then split, by screening, to 14 35 and 35 150 mesh fractions, which are subjected separately to a second cationic flotation step using high-quality cationic reagent to reject the remaining coarse silica and achieve a final phosphate concentrate assaying at least 30% P2O5 with superior economic yield. In this technique, corn starch

Table 3 Hamrawain Phosphate Sample (the feed assayed 19.0% P2O5) (El-gillani and Abouzeid, 1993) Product Carbonate float Phosphate conc. Carbonate float Phosphate conc. Carbonate float Phosphate conc. Carbonate float Phosphate conc. Carbonate float Phosphate conc. Carbonate float Phosphate conc. P2O5% 10.51 27.62 8.49 30.0 7.11 31.65 7.44 31.84 7.63 30.81 7.45 31.91 Insoluble, % 6.0 3.38 5.47 3.22 5.94 3.17 5.5 2.45 6.41 2.77 5.87 2.8 LOI, % 31.25 11.55 33.47 9.8 34.6 7.8 35.5 7.54 34.06 8.21 34.3 7.45 P2O5 Rec., % 68.51 77.96 79.91 79.7 77.62 78.13 Modifying reagents a 2.2 kg/t H2 SO4 a 4.0 kg/t H2 C2 O4 a 2.96 kg/t H3 PO4 a 9.6 kg/t KH2 PO4 b 3.2 kg/t KH2 PO4 b 1.5 kg/t H2 SO4 2.4 kg/t KH2 PO4 b 2.0 kg/t H2 C2 O4

Phosphate depression and pH controller, collector is 0.8 kg/t in all experiments. a Reagent was added during conditioning and flotation. b KH2 PO4 was added before conditioning and pH was adjusted by H2 SO4 and/or H2 C2 O4.

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

71

proved to be both efficient and cost effective as a phosphate depressant (Snow et al., 1996). A similar technique was reported by Lukomona et al. (2005) where they treated an apatite ore assaying 22.7% P2O5, 22.8% SiO2 (silica and silicates), 7.0% Fe2 O3 (mainly as magnetite), and 4.0% Al2 O2 by a single step cationic flotation process. They used petroleum sulfonate as a cationic collector at an optimum pH of 9.5. The produced phosphate concentrate assayed 40.8% P2O5 and 1.65% SiO2 at a recovery of 50.2% P2O5 (Lukomona et al., 2005). On the other hand, Comonco Company (1974) registered a Canadian Patent for the all-anionic phosphate flotation technique. In this process, the ground feed is subjected to an anionic flotation step to reduce the siliceous gangue constituents. The froth floated product is treated with phosphate and sulfate salts as depressants for gypsum, and then subjected to another anionic flotation step (Comonco, 1974; Johnston and Leja, 1978). The sulfate depressant should be a salt, but not sulfuric acid (Comonco, 1974). Similarly, an Indian siliceous calcareous phosphate ore was beneficiated through a two-step anionic flotation scheme. In the first step, phosphate and calcite are floated and silica is depressed. In the second step, calcite is floated and phosphate is depressed. The reagent scheme consisted of sodium oleate, sodium silicate, phosphoric acid, and sulfuric acid. A phosphate concentrate assaying 27.2% P2O5 at a recovery of 73% was obtained from a feed assaying 12.5% P2O5, 26.5% SiO2, and 32% CaCO3 (Prasad et al., 1995). Non-ionic polymers were tested for selective flotation of phosphate rocks. Miller (2001) added polyethylene oxide, PEO, (molecular weight between 1000 and 8000) to the anionic collector to float Florida phosphate ore. It was demonstrated that, for the same phosphate recovery, the collector amount was reduced in the presence of PEO (about 10% of the collector weight) down to one-half of the amount required without the addition of PEO. When the same amount of collector was added in presence and in absence of PEO, the phosphate recovery was 1030% more with PEO addition. This type of polymer is more effective in the flotation of coarse phosphate than in the flotation of fine phosphates. The improved phosphate recovery with PEO addition was attributed to the improved froth stability and improved dispersion/ adsorption of fatty acid due to the apparent interaction between PEO and the fatty acid (Miller, 2001; Lu et al., 1999). Pinto et al. (1991) reported that a sulfosuccinate and a sulfosuccinamate proved to be successful in phosphate flotation as new collectors. It was observed that the collector ability as well as the flotation rate of each reagent is strongly dependent upon the mineral size range. These

collectors displayed higher flotation rates for coarse particles. Synergistic effects were also apparent when a conventional collector (sodium oleate or rice bran oil) was mixed to one of these new collectors. Another phosphate flotation collector was patented by Yu and Xue (1996). This collector comprises a mixture of: tall oil, fuel oil, N-substituted-N-mixed alkyoxy-propylmaleimic acid derivative, fatty acid sulfonated, alcohol ether sulfate, and alkyl alcohol sulfate. It was used successfully to float phosphate minerals. The collector is added to the aqueous slurry of the phosphate ore and conditioned in a conventional manner. The pH is adjusted with alkaline modifier at pH 7.011.5. The conditioned pulp is floated in the traditional way (Yu and Xue, 1996). An alcoholic solution of alkyl hydroxamic acid collector was also used for phosphate flotation (Miller, 2002). It functions as a wetting/spreading reagent in a similar manner to the oily/ water insoluble collectors, but with greater selectivity. High solids conditioning is preferred to create stable buoyant aeroflocs. When the alkyle hydroxamic acid is used as a collector, a clean phosphate concentrate is obtained in a single-stage flotation operation. In the case of coarse phosphate feed (15% P2O5), the concentrate grade reached 34% P2O5 with 93% recovery, while a concentrate assaying 29.6% P2O5 with a recovery of 96.0% when fine feed was used (Miller, 2002). There are three basic groups which are utilized in the application of frothers namely: alcohol, polyglycol, and polyglycol esters. Each chemical group provides particular characteristics which affect load, bubble diameter, water retention, air dispersion, and kinetics of froth formation. A frother may comprise a single chemical group or a blend of more than one group depending on the particular ore to be treated and the cell environment (Riggs, 1989). Specific attention was given to the development of new frother and to improving the frothing characteristics of the existing ones. Moudgil and Gupta (1989) claimed that the phosphate flotation recovery was increased by 3% when they used specific types of frothers. Addition of fine particles was also found to enhance flotation of coarse fractions under given conditions. The grade of the product with and without frothers remained the same. Change in froth stability and viscosity are believed to be the primary reason for enhanced recovery (Moudgil and Gupta, 1989). A number of different chemicals were used to depress dolomite in the flotation of apatite and collophane. Carboxymethyle cellulose, citric acid, and naphthyle anthyle sulfonates were found to be good depressants for dolomite with the cellulose as the best among them. None of these depressants significantly affected the flotation of apatite or collophane (Zheng and Smith, 1997).

72

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

On the other hand, the study of activation of collophane by fluoride ions using Hallimond tube flotation tests was carried out by Miller et al. (1987). The results revealed that efficient oleate flotation of collophane can be achieved by fluoride activation. At low fluoride concentration, fluoroapatite type compound may form on the collophane surface. At high fluoride concentration, calcium fluoride appears to form on the collophane surface. As a matter of fact, the efforts spent in solving the problems of separating carbonates from phosphate ores, particularly in the case of sedimentary phosphate ores, are tremendous. Most of these efforts are documented by the Florida Institute of Phosphate Research, FIPR (Zhang et al., 1999). This extremely important publication is a collection of abstracts of most of the published material on phosphate concentration all over the world since 1930. 5. Thermal treatment of phosphate ores More than 10% of the world marketable phosphates are produced by calcination. Traditionally, heat treatment of phosphate ores is defined as heating up the ore to a certain temperature to obtain a product with specific properties. The main objectives behind heating up the ore are: i) Removing water ii) Removing organic matter iii) Dissociation of carbonates iv) Removal of fluorine 120150 C 650750 C Drying Calcination

the organic matter and to oxidize the residual carbon. In this case, exothermic reactions (as a result of burning organic matter and oxidizing residual carbon) take place at this level of temperature. The organic matter in the ore contributes appreciable amounts of the heat requirement for calcination (Bandyopadhyay, 2006). The other type of calcination of phosphate ores is carried out for thermal dissociation of carbonates. The calcination temperature in this case ranges from 850 1000 C (Bandyopadhyay, 2006). The dissociation of carbonates is a highly endothermic reaction which requires large fuel consumption. The suggested dissociation reaction goes as follows (Kunii and Levenspiel, 1969): CaCO3 Heat CaO CO2 42:9kcal=mol Defluorination is the process of obtaining fluorinefree highly purified forms of phosphorus-containing products to be used as supplements in food ingredients, livestock and poultry feeds, with phosphorus in water soluble forms. These products are essential for the health of the animals. Since fluorine is toxic to animals, it has to be removed almost completely by defluorinating the phosphate ores. Defluorinated phosphates are produced by adding defluorinating agents, such as phosphoric acid and soda compounds, in controlled amounts to phosphate rock and calcining the mix at temperatures ranging from about 1370 C to 1510 C conditions that drive off the fluorine and converts the unavailable phosphate to the biologically available form. The process transforms phosphate mineral to the defluorinated phosphate with a P/F ratio of more than 100. Calcined product is cooled and screened to defluorinated phosphate rock (DFP) or tri-calcium phosphate (TCP) (Bandyopadhyay, 1999). Because it has a low enough fluoride to phosphorus ratio, and the phosphorus is bio-available, it is suitable as an animal feed supplement. The product is essentially a solid solution of calcium phosphates as well as calcium silicate. The principle reaction is shown below. No attempt has been made to balance the equation, which comprises a far more complex set of chemical reactions (Coffey et al., 1994): 3Ca3 PO4 d CaF2 s SiO2 s NaOHaq H3 PO4 aq 3Ca3 PO4 2 d Na3 PO4 d Ca2 SiO4 s 2HFg

8501000 C Calcination up to 1350 C Defluorination

Drying, which is a common step in the wet processing of phosphates, is categorized under thermal treatment, but it is not included under the term calcination. Almost all these heat treatment operations are preceded with ore preparation such as: crushing, grinding, sizing and/or classification, and possibly flotation, to get rid of some of the liberated impurities at one or more of these stages. Calcination of phosphate ores is necessary for obtaining a phosphate product suitable for the production of light green phosphoric acid, which is in turn suitable for the production of pure, edible, super phosphoric acid. This last product is used in the preparation of pure chemical reagents, food, livestock, soft drinks, and other pharmaceutical preparations. There are two types of calcination operations. One is associated with organic matter. In this operation, the ore is heated up to 650750 C to burn

Fig. 4 shows a schematic flow sheet for the defluorination process of the phosphate rock (Bandyopadhyay, 1999).

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

73

5.1. Energy consumption in thermal treatment of phosphate ores The amount of energy for mining and physical beneficiation of phosphate ores to produce one ton of commercial rock phosphate (32% P2O5) is approximately 0.9 GJ/ton of rock phosphate (about 2.9 GJ/ton P2O5). However, if the rock is calcined, the total energy requirement is doubled (Ullmann's Encyclopedia, 1991). Considering all the major factors consuming heat in the calcination process, the heat required to produce 1 kg of calcined phosphate from a raw feed containing about 50% calcium carbonate and 10% moisture was found to be in the range of 800 kcal/kg of calcined product. This amount of heat is equivalent to about 0.15 kg of coal of calorific value in the range of 7000 kcal/kg of coal. This amounts to 3.6 GJ/ton of calcined product (Abouzeid, 2006). It was reported by Bandyopadhyay (2006) that the calcination process involves drying of beneficiated phosphate rock (from 1015% moisture down to 1 3% moisture) followed by calcination at temperature of about 800 C. The data provided by Bandyopadhyay was

basically for the treatment of the North Carolina phosphate ores. This type of ore contains organic matter as the main impurity in the ore. For this reason the cost for the calcination process was considerably low. He reported US$ 1.68/short ton for drying and US$ 3.49/ short ton for drying and calcination at 800 C based on the 1999 cost level. In this case, the energy consumption for calcination was very low due to the presence of organic matter which supplied most of the calcination heat during burning. The energy cost for defluorination of phosphates at 1350 C was in the range of US$ 36.2/ short tone of product (Bandyopadhyay, 2006). It is thought that the feed to the defluorination plant was also a North Carolina phosphate rock, which means that for defluorination of a calcareous ore, with little or no organic matter, the energy required will be higher. 5.2. Limitations and benefits of high temperature treatment of phosphate ores There are several factors that are traditionally known as drawbacks for the calcination of phosphate ores

Fig. 4. Defluorination process of phosphate rocks (Bandyopadhyay, 1999).

74

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

(Abouzeid et al., 1996; Orphy et al., 1969). Among these factors are: 1-. High capital cost of calcination plants. 2-. Energy extensive operation. 3-. Low reactivity products. 4-. It usually affects gypsum filtration rate, adversely (Bandyopadhyay, 1999). On the other hand, there are factors that are in favor of phosphate calcination. Among these factors are: 1-. Upgrading of the ore, reduces the tonnages to be treated for the same P2O5 units, and, hence reduces costs for transportation and fertilizer and/or phosphoric acid production. 2-.Reduces the amount of sulfuric acid required for acidulation of the ore in the subsequent chemical processing. 3-. In most cases, the drawback of low reactivity of the calcined product could be avoided by controlling calcination temperature below 800 C when the main impurities are organic matters. In this type of phosphate ore, with high organic matter content, burning of organic matter is exothermic (evolves heat) which saves energy. 4-. Eliminating the foaming phenomenon during acid attack of the calcined phosphate products. Elimination of the foaming solves many technical problems caused by the presence of organic matter and/or carbonate minerals, and saves the costs of the defoaming reagents. 5-. The capital costs of the calcination plants are comparable with the drying units which are commonly included in the wet beneficiation flow sheets. 6-. The special products obtained by high temperature treatment (super phosphoric acid and defluorinated phosphates) are highly demanded and highly priced. Any way, the final decision for obtaining the required phosphate product will be taken based on thorough analysis of the economic and technical assessment of the process and on the available type of ore that will be used as the initial feed. 5.3. Types of thermal treatment units There are various types of units for calcination of phosphate ores. These include: 1-. Vertical shafts (Emich, 1984; Disanto, 1985; Ball et al., 1973). 2-. Fluidized bed reactors (Kunii and Levenspiel, 1969; Priestly, 1967; Perry et al., 1963).

3-. Rotary kilns (Emich, 1984; Disanto, 1985). 4-. Traveling grate kilns (Ball et al., 1973). 5-. Traveling grate-rotary kiln systems (Emich, 1984; Ball et al., 1973). 6-. Flash calciner (Emich, 1984; Blazy and Bouhaouss, 2005; Henin and Lectard, 1983; Henin and Pinoncely, 1986; Disanto, 1985; Ersahan et al., 1994; Bridson et al., 1985). Readers may be referred to the cited references for detailed designs, and operating techniques, and advantages of each of the above units. 5.4. Industrial calcination plants Worldwide phosphate calcining capacity in 1981 was about 14 Mt/a. About 75% of world calcination is treated in Dorr-Oliver FlouSolids systems. Allowing 85% equipment availability, the yearly design capacity was 8.9 Mt/a. Actual production is probably 120% of design capacity, or 14.3 Mt/a for all calcination installations. These installations are designed to reduce the organic content of phosphate rock. However, in Algeria, the calciners, which are operated at higher temperatures, are used to drive off CO2 from carbonates. Table 4 gives the most important world phosphate calcination plants for the removal of organic matter as well as calcite (Emich, 1984). 5.5. Effect of substituting carbonates on the characteristics of carbonate-apatite The apatite in phosphorites and rock phosphates usually are poorly crystallized (Zafar et al., 1995; Abouzeid et al., 1980). Their compositions differ considerably from those of pure fluorapatite or hydroxy-apatite. Their X-ray diffraction patterns are typically apatite with slight shifts that show changes in the cell parameters (Abouzeid et al., 1980). As already mentioned, most sedimentary phosphor ites contain considerable amounts of carbonate, CO3 2, (Palache et al., 1951), up to 6.3% by weight of the phosphate mineral, and are often considered to be
Table 4 World phosphate calcining capacity in 1981 (Emich 1984) Installation Location Djebel Onk OCP Beker Simplot Stauffer Texas Gulf Total Algeria Morocco Idaho Idaho Wyoming North Carolina Calcination units Design capacity, t/d 3 2 4 4 3 6 22 3000 3700 5830 5000 2000 9000 28,530

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

75

carbonate-apatite or francolites (McClellan and Lehr, 1969). McConnell and Gruner (1940) found small distinct changes in the intensities of certain francolite reflexions from those of apatite reflexions. However, Hendricks and Hill (1950) reported that X-ray goniometer photographs of francolite crystals are indistinguishable from those of apatite crystals, although theoretically they should. Altschuler et al. (1952) presented X-ray evidence for carbonate flourapatite as a distinct variety of apatite rather than a mixture of a carbonate phase and fluorapatite. Silverman et al. (1952) investigated the differences in X-ray patterns of carbonate-fluorapatite, hydroxy-apatite, and fluorapatie. They detected the presence of carbonate in the structure of carbonate-apatite and differentiated it from that of fluorapatite. Maslennikov and Kavitskaya (1956) and Alden and Lindgvist (1964) recognized that the a-axis in apatite containing both fluoride and carbonate ions would be short (9.32 to 9.35 ) compared to the a-axis of igneous and metamorphic apatite (9.38 to 9.40 ) that contain fluoride ions with almost no carbonate ions. Such variations in the a-dimension indicate that the excess CaF2 or CaCO3 cannot be simply adsorbed on the surface of the apatite crystals, but it may correspond to some regular type of overgrowth that leads to a series of compounds with the following end-members general formulae (McClellan and Lehr, 1969): Ca10 PO4 6 F2 Ca10ab Naa Mgb PO4 6x CO3 x F2 d Fy Electroneutrality of the above equation is preserved when a = x y. Experimentally it was found that y = 0.4x where x varies from near zero to 1.5 and b = 0.4a. These end-members correspond to the chemical composition shown in Table 5. The maximum degree of substitution was found for carbonate-francolite type apatite when x / (6 x) 0.3 which is the case in marine phosphate sediments. The fluorapatite (x = 0), which is the main phosphate mineral in igneous and metamorphic rocks, has the highest concentration of P2O5 (42.2%). The most significant feature revealed is that the relative increase in amounts of F, Mg, Na, and CO2 is accompanied by a decrease in P2O5 content and nearly constant CaO content (Palache et al., 1951). Apatite prepared with increasing contents of carbonate and decreasing content of phosphate exhibit proportional changes in the dimensions of the a-axis. As no phase other than apatite is detectable from the X-ray patterns, it was concluded that the carbonate is a substitute for the phosphate within the structure. Smith and Lehr (1966) indicated that there is a reasonable correlation between the total fluorine content and the carbonate content and the

Table 5 Comparison between chemical constituents of fluorapatite and francolite (Abouzeid et al., 1980) Constituent, % CaO P2O5 CO2 F Na2O MgO F1uorapatite (x = 0) 55.6 42.2 0.0 3.77 0.0 0.0 Francolite (x / (6 x) 0.3) 55.1 34.0 6.3 5.04 1.4 0.7

length of the a-axis. The increase in carbonate content in the lattice structure would cause a decrease in the a-axis of the unit cell. McClellan and Lehr (1969) confirmed the findings of Lehr (1964) and showed that there is a progressive change in the unit cell dimension as (CO3 + F) replaces (PO4) and that the minimum length of the a-axis corresponds to a high substitution of carbonate (Lehr and McClellan, 1974). The increase in carbonate and fluorine substitution will lead to a decrease in crystallite size. Crystallite sizes range from microscopically coarse crystals for igneous and metamorphic apatite to submicroscopic crystallites, only a few hundred Angstrom units, in the most highly substituted sedimentary apatite. Le Geros et al. (1967) concluded that the presence of carbonate causes diminution in size and a change in the shape of apatite crystallites. Carbonate-free apatites are thin, elongated crystallites while the apatite rich in carbonate are more spheroidal crystallites with a length-to-thickness ratio approaching unity. They also concluded that the changes in size and shape possibly were due to the formation of weak bonds by carbonates, and that the growth of a crystal is less in the general direction of the formation of weak bonds than in the general direction of the formation of strong bonds. Cuthbert and Rowland (1947) studied the differential thermal analysis of ten different carbonate minerals and found that they dissociated at different temperatures. All the carbonates showed endothermic reactions caused by dissociation, but siderite showed an exothermic reaction caused by the decomposition of FeCO3, which gives FeO which is immediately oxidized to Fe2O3 (Cuthbert and Rowland, 1947; Plotzki, 1937). The differential thermal curve of calcite showed an intense broad endothermic reaction starting at about 625 C and ending at about 893 C with a peak at 840 C. The aragonite curve also showed a broad endothermic reaction starting at about 600 C. In the same direction, Faust (1944) made several studies using differential thermal analysis to differentiate between magnesite and dolomite and between calcite and aragonite (Faust, 1950). Silverman et al. (1952) detected the carbonate associated as calcite from bound carbonate

76

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

Fig. 5. Weight loss as a function of calcinations temperature for two Jordanian calcareous phosphate samples (an original sample and a calcite-free sample) (Abouzeid et al., 1980; El-Jallad, 1977).

in carbonate-bearing apatite by means of DTA analyses. Confirmation of these findings was observed by El-Jallad (1977) and Abouzeid et al. (1980) for two francolite phosphate samples; one includes associated calcite and the other is calcite-free sample. Fig. 5 shows that the rate of dissociation of carbonate-apatite (the calcite-free sample) is higher than that of the sample containing associated calcite. This means that the release of CO2 from carbonate-apatite is easier than from calcite probably due to internal stresses inside the carbonate-apatite structure. It should be noted that the two samples are of the same ore sample with the calcite-free sample prepared by dissolution of calcite using Silverman solution (Triammonium citrate solution) (Ames, 1959). Phosphate ores of high carbonate content could be beneficiated by heating the rock to temperatures between 600 and 950 C. This removes free and combined moisture and most of the carbon dioxide (Khalifa et al., 1972; Stern, 1965). Clay, iron and aluminum oxides, if present in the phosphate ores, are very harmful impurities in the calcination process. If the ore is not washed, the clay will be hardened by dehydration, and sometimes it may form silicates with magnesium and calcium oxides that are responsible for some technical troubles and clinker formation in the kiln especially at calcination temperatures about 1100 C (Khalifa et al., 1972; Rudis, 1967). Washing the ore before calcination, therefore, is very important. The calcination temperature plays a decisive role in determining the time required for hydrating the lime. Petrographic studies showed that phosphate samples heated at 1200 C contained free CaO, some undecomposed CaCO3, slaked

CaO, and clinker material in a glass phase, and, in some cases, gypsum. The undecomposed CaCO3 was found in single grains or mixed with aggregates of CaO (Khalifa et al., 1972; Nikitin et al., 1966). 5.6. Characteristics of calcined phosphate product The thermograph of the material to be calcined as a function of temperature is different from one type of phosphate ore to the other depending whether the ore contains organic matter, carbonate minerals, or both of them as the main gangue material. If organic matter is dominating, the reaction will be exothermic, and less energy is required to initiate the reaction (Fig. 6) (Blazy and Bouhaouss, 2005). It should be noted that there is an exothermic reaction between 625 C and 800 C probably due to recrystallization of the carbonate-apatite constituents. With carbonates, e.g. about 50% calcite, as gangue minerals, with little or no organic matter, the reactions will be endothermic needing around 400 kcal/kg of heat energy for dissociation of carbonates (Abouzeid, 2006), Fig. 7. This is a large amount of heat energy to be supplied to the system for calcination. When both organic matter and carbonate minerals exist, the thermograph reflects exothermic peaks, endothermic peaks and interference regions where both exothermic and endothermic reactions overlap (El-Jallad, 1977). In this case the amount of heat energy needed for calcination will depend on the relative existence of the two constituents, organic material and carbonate minerals. Fig. 8 shows the differential thermograph for a phosphate ore containing carbonates and organic matter as

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

77

Fig. 6. Differential thermal analysis for a Moroccan phosphate ore, rich in organic matter (Blazy and Bouhaouss, 2005).

the main gangue constituents (El-Jallad, 1977). In this figure, there is a series of weak endothermic reactions as a result of moisture evaporation and removal of water of crystallization with a possibility of overlapping an exothermic effect with a peak at 470 C due to burning of organic matter. There is also a strong exothermic reaction with a peak at 750 C, probably due to recrystallization of the carbonate-apatite, followed by a strong endothermic reaction caused by the dissociation of calcite. Differential thermal analyses were also done on different phosphate samples from different localities in Jordan by Rosch and Saadi (1970). They noticed that there was an endothermic

moisture peak between 100 and 200 C, two strong exothermic peaks at temperatures between 300 to 400 C and 730 to 770 C, and strong endothermic carbonate doublet peaks between 795 to 815 C and 830 to 890 C. 6. Effect of calcination on physical properties of the phosphate rock 6.1. Unit cell dimensions and crystal size Chemical analysis of calcined phosphate showed higher P2O5 grade in the phosphate product after

Fig. 7. Differential thermal analysis for the thermal dissociation of calcite (El-Jallad, 1977).

78

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

Fig. 8. Differential thermal analysis for a Jordanian phosphate ore containing organic matter and calcareous gangue material (El-Jallad, 1977).

calcinations (Khalifa et al., 1972; Flaszenberg et al., 1962; Hegner and Pacl, 1974; Saadi, 1969; El-Jallad, 1968; Masson, 1959; Alekseev and Shakmatova, 1964). Stolovitskaya (1966) pointed out that the degree of recrystallization of CaO depends mainly on the calcination temperature. Decomposition of the adsorbed and combined water, burning of organic matter and most of the combined carbon dioxide increase the phosphate grade in the calcined product. Freeman et al. (1964) Ando and Matsono (1966) found that the a-axis of the unit cell started increasing when the carbonate-apatite was heated above 500 C. When heated above 900 C, it reached the value of the aaxis of fluorapatite. The X-ray diffraction peaks of fluorapatite became more distinct, sharper, and stronger by calcination above 700 C. They attributed this behavior to crystal growth of apatite through the combination of fine crystals of the initial phosphate forming larger crystals, which upon calcination at about 950 C, resembled Kola apatite. Smith and Lehr (1966) found that the major effect of calcining carbonate-apatite above 800 C was the evolution of carbon dioxide with a consequent increase in crystal size and the recrystallization of stoichiometric fluorapatite and the segregation of isotropic CaO and CaF2. They also found that the loss of carbon dioxide of the carbonate-apatite, by calcination, caused an increase in the unit cell a-axis. Abouzeid et al. (1980) obtained similar results concerning the increase in the a-axis and the growth of crystal size as the calcination temperature increases. The value of the a-axis of the unit cell increased from 3.335 at room temperature to 3.364 after calcination at 950 C (Fig. 9). The crystal

size of the phosphate mineral was increased from 500 at room temperature to 2200 after calcination at 950 C (Fig. 10). It can be noted that the presence of calcite and/or calcium oxide does not affect the values of a-axis or the crystal size at all calcination temperatures. Maslennikov and Kavitskaya (1956) noticed an increase in the length of the unit cell a-axis of the carbonate-apatite after calcination. This increase was due to the loss of carbon dioxide in the carbonateapatite. Lately, Blazy and Bouhaouss (2005) reported that the value of the a-axis of the unit cell continuously increased with temperature until it reach 9.357 at 950 C when calcination was carried out for four stages in a flash calcination system. 6.2. Surface area, porosity, and material density 6.2.1. Surface area Freeman et al. (1964) reported that the specific surface area of the calcined product of phosphate rock decreased with increasing calcination temperature. They carried out their tests on two different phosphate samples from Morocco and Utah. There was a sharp decrease in specific surface area with increasing temperature in both samples between 500 C and 800 C, with the Morocco sample more pronounced in the density decrease. They attributed this remarkable change in surface area to the aggregation of elementary fine grains in this temperature range suggesting that consolidation of these very small grains occurs at these temperatures. It may also be possible that the destruction of the pore structure of the rock in this temperature range contributes to the drastic

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

79

Fig. 9. The unit cell dimension (a-axis) as a function of calcinations temperature for two Jordanian phosphate samples (an original sample and a calcite-free sample) (Abouzeid et al., 1980; El-Jallad, 1977).

decrease in specific surface area of the sample as will be discussed later. 6.2.2. Porosity The porosity of the phosphate rock sometimes increases with increasing temperature up to about 700 C and then decreases from this level on as in the case of the Utah phosphate rock (Freeman et al., 1964). In some other phosphates, such as the Morocco phosphate, the porosity gradually decreased with increasing temperature and then suddenly fell to minimum at about 700 C. The porosity trends suggest the destruction of the pore systems in both

samples for both the macro pores and Micro pores, but particularly the fine pores. For the Morocco sample the destruction of pores was quite severe which may be reflected on the reactivity of the calcined product. It is worthy to mention that in the case of the Utah sample, the gradual increase of porosity in the lower range of temperature (below 500 C) may be explained on the basis of volatilization of organic matter (Freeman et al., 1964). 6.2.3. Density It was observed that there is little change in particle density below 750 C. Above this temperature, some

Fig. 10. The mean crystal size of Jordanian phosphate samples as a function of calcinations temperature (Abouzeid et al., 1980; El-Jallad, 1977).

80

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

changes in density occur, where the density slightly increases with increasing calcination temperature. The particle density roughly reflects the internal collapse of the particle porosity between temperatures, of 650 C to 750 C. The true density is almost not affected by the calcination temperature except at relatively high temperatures where crystal transition may take place and cause the density to be lower (Freeman et al., 1964). 6.3. Effect of calcination on chemical reactivity of the calcined product The chemical reactivity was found to be a reasonable measure of the suitability of the phosphate rock for various purposes. The solubility behavior of phosphorus in mild solvents such as citric acid, neutral ammonium citrate, and alkaline citrate provides a very good indication of relative reactivity. The amount of soluble phosphate extracted from a certain quantity of phosphate rock in a certain volume of solvent under specific conditions is called the relative solubility or reactivity of the rock. Diluted mineral acids may also be used for measuring the relative reactivity of the phosphate rock. The reactivity of a phosphate rock is an inherent property of the mineral that influences its rate of reaction with acids (Freeman et al., 1964). It seems also that the occurrence of the phosphate minerals as well as their associated constituents and the extent of the physical and chemical changes as a function of calcination temperature are highly effective factors on the rate of dissolution of phosphorus in various acids or salts references. The

phosphate rock reactivity is a strong function of the unit cell dimension, a-value, which is a reflection of the phosphate mineral present. Fig. 11 shows the trend of the chemical reactivity of some phosphate samples as a function of the unit cell dimension (Straaten, 2002; Abouzeid et al., 1980). This figure demonstrates that the reactivity of phosphate rocks decreases with increasing the a-value of the unit cell. The chemical reactivity of the phosphate rock with a solution of 2% citric acid decreased slightly upon calcination at 500 C because of the decrease in porosity. It also decreased sharply when the phosphate rock was calcined above 900 C as a result of changes in unit cell dimensions (recrystallization of the phosphate minerals), crystal growth, and decrease in porosity and surface area as mentioned before. Treatment of the phosphate rock in a 0.25 N hydrochloric acid showed a slight decrease in reactivity after calcination. This decrease, however, was not significant compared to the original uncalcined phosphate rock. The rate of decomposition of phosphates calcined above 900 C in a mixture of sulfuric and phosphoric acid showed a remarkable decrease in their chemical reactivity. Lehr (1964) and Smith and Lehr (1966) studied the solubility behavior (chemical reactivity) of 17 phosphate rocks before and after calcination by using neutral ammonium citrate solution for the reactivity measurements. They found that the chemical reactivity decreased slightly above 600 C owing to the annealing of phosphate crystals. However, there was a sharp drop in the chemical reactivity of the phosphate rock after calcination above 750 C caused by

Fig. 11. The solubility index of phosphate rocks as a function of the unit cell dimension, a-value (Straaten, 2002; Abouzeid et al., 1980).

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

81

Table 6 Percent relative reactivity of some phosphate rocks of different geographic locations, using neutral ammonium citrate solutions (Abouzeid et al., 1980; Freeman et al., 1964) Geographic location of the deposit Jordanian phosphate Jordanian phosphate Morocco phosphate Utah phosphate Peru phosphate Florida phosphate Reactivity index of uncalcined phosphate rock 24 43 87 60 75 52 Calcination temp. C 950 950 910 775 900 815 Percent of reactivity index of uncalcined rock 12.5 41.9 53.4 89.7 93.8 95.0 Remarks CO2(total) = 19.1% CO2(apatite) = 6.1% Calcite-free CO2(apatite) = 6.1% CO2(total) = 5.2% CO2(apatite) = 4.8% CO2(total) = 2.7% CO2(apatite) = l.6% CO2(total) = 4.6% CO2(apatite)4.2% CO2(total) = 2.4% CO2(apatite) = 2.2%

recrystalization. It reached its lowest value when the ore is calcined at about 950 C owing to phosphate mineral transformation to fluorapatite and extensive crystal growth. Hegner and Pacl (1974) found that the chemical reactivity of calcined phosphates decreased as the calcination temperature increased and reached its lowest value at 900 C. They used in their studies 2% citric acid, 0.25 N nitric acid, 0.25 N sulfuric acid, and 0.25 N phosphoric acid. They attributed the drop in chemical reactivity of the calcined phosphate, at about 900 C to an alteration of carbonate-apatite to fluorapatie and to a decrease in the specific surface area. El-Jallad (1977) found that this decrease in chemical reactivity could be caused by changes in crystal structure, crystal size, and the presence of CaCO3 and CaO in the product. It has been reported that phosphate rocks from different geographic locations respond differently, from the physical and chemical aspects, to calcination. Changes in some cases are negligible, and in some other cases are highly remarkable. Table 6 gives some examples of changes in reactivity indices of some phosphate samples from different geographic locations (Bandyopadhyay, 2006; Abouzeid et al., 1980; Freeman et al., 1964). It is obvious from Table 6 that the reactivity index is different for different phosphate deposits differing in geographic locations. It also shows that the drop in reactivity index of the calcined product can be as high as 88% of the uncalcined rock and as low as 5%. This may be due to the limited amount of dissolving acid, and the relative amount of calcium carbonate and calcium oxide which affect the reactivity index to a noticeable extent. It should be mentioned that in the case of calcined calcareous ores, some chemicals may be added to enhance the dissolution of calcium and magnesium hydroxides during the washing stage. The most common of

these chemicals is ammonium chloride, NH4Cl. A concentrate assaying 32% P2O5 and 1.3% CO2 was obtained from a feed assaying 21.3% P2O5 and 19.5% CO2 at a recovery of 98.2% after washing the concentrate three times with 1% solution NH4Cl. When water, without addition of chemicals, was used in washing the above concentrate, the assay was 27.5% P2O5 (Ahmed, 1990). Orphy et al. (1969) calcined raw feeds assaying 24.8 30.5% P2O5. When they washed the calcined product with NH4Cl solution (20 kg/ton of ore), they obtained concentrates assaying 35.437.1% P2O5, at P2O5 recoveries of 91.498.1%. 7. Summary Techniques for beneficiation of phosphate rocks depend on the type of the phosphate deposit, the phosphate minerals existing in the ore as well as the associated non-phosphatic material. Simple beneficiation techniques such as crushing and screening or scrubbing and desliming are successfully used when the main gangue material is hard siliceous material, silicates and/or clayey material. Ores containing quartz, chalcedony or different forms of silica, may be upgraded using gravity and/or electrostatic techniques. However, the use of these techniques for large scale production is limited due to their limited capacities. The main conventional concentrating technique for such ores is flotation. Flotation in this case is carried out either in one stage or two-stage operation. The one stage technique is mainly to float the phosphate minerals using oleic acid or sodium oleate with the addition of fuel oil (kerosene) as collectors at pH = 8. The two-stage process includes, in addition to the previous step, decreasing the value of pH to 4,

82

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984

skimming the desorbed oleic acid, and adding cationic collector to float the trapped silica. This technique produces a clean high-grade phosphate product. Igneous phosphate rocks are concentrated mainly by anionic flotation of phosphates. The flow sheet for upgrading this type of ore may include intermediate steps depending on the type of gangue minerals associated with the phosphate minerals. If sulfides exist, they are removed by flotation. Magnetic minerals are removed by magnetic separation, and clays and fine impurities are removed by attrition scrubbing and desliming. The concentrate from these types of rocks is generally of high grade (4042% P2O5) and is used for the production of edible phosphoric acid. Phosphate ores containing high content of organic material as the main ingredients are processed by calcination at about 500 to 750 C to burn out the organic material. Sometimes, calcination temperature is higher than this range, which may be as high as 800 C to burn the residual carbon. At this low range of temperature, physical and chemical properties of the phosphate minerals are not significantly affected and hence, the reactivity index of the phosphate is not remarkably affected. The reactions taking place as a result of raising temperature are exothermic reactions which does not need large additional amount of energy. The calcined product in this case is suitable for the manufacture of clear green phosphoric acid which is used as a source for producing super phosphoric acid and other high-grade chemical reagents. The calcareous sedimentary phosphate ores form the main challenging problem in phosphate processing. Electrostatic and gravity separation are not yet up to the large scale production techniques. Extensive efforts have been spent to use flotation, which is the most widely technique used in processing most of the ores with different modifications and developments, for upgrading the calcareous sedimentary phosphate ores. On the laboratory scale, some of these approaches were successful, but they are not tested in pilot or industrial scales. The rest of these efforts were neither practical nor successful. To utilize these natural phosphate deposits, some of them are being calcined and/or defluorinated. Although these products are of inferior quality with respect to the chemical reactivity and extensive amounts of energy are used, they are almost pure products suitable for producing high-grade chemical reagents and/or defluorinated phosphates for animal feed stocks. These calcines are of high quality and are sold as expensive commodities for special purposes. In the case of using these products for producing phosphoric acids defrothing reagents are not needed, and filtration of residual solids are made easier.

Researches for economic and technical utilization of the huge calcareous phosphate deposits around the world are urgently needed. Thorough understanding of the thermodynamics of the carbonate/phosphate/water systems is of high priority. Development of suitable flotation reagents that are more selective to phosphate or carbonate minerals is another incentive to apply flotation in upgrading the calcareous sedimentary phosphate rocks. Carefully adjusting the suspension environment (pH, ionic strength, etc.) to differentiate between surfaces of phosphates and carbonates for better particle/collector and particle/bubble interactions should be considered. The possibility of using electrostatic and/or gravity separators for concentrating phosphates on industrial scale may help solving the problem. Finally, the flash calcination is a promising technique for economic calcination of phosphate ores when it is scaled up to produce large tonnages of phosphate calcines. The product obtained by this technique is of superior quality than that produced by the other conventional calcination techniques. The laboratory experimental results show that effective reduction in carbonate content of phosphate ores can be achieved by organic acid leaching of the calcined calcareous phosphate ores. If economy favors this technique, and optimum conditions on a large scale pilot testing confirm the successful results obtained in the laboratory, it should be seriously considered, particularly because organic acids are easily degradable, less environmental hazardous, and can be regenerated. Acknowledgements The author would like to thank Dr. Bhaskar Bandyonadhyay, consultant, Professor Peter van Straaten from the University of Guelph, Canada, Miss Karen Stewart, Librarian at Florida Institute of Phosphate Research, and the phosphate Staff at the United States Geological Survey for their close contact and appreciable help. They provided the author, generously, with the necessary information regarding the subject. The author's appreciation is also extended to Miss Eanass Abouzeid for her patience in typing the draft of the paper and reproducing the attached figures. References
Abdel Moneim, H.M., 1971. Geological investigations and physical beneficiation of some low grade phosphate rocks in Egypt, M. Sc. Thesis, Ein- Shams University. Abouzeid, A.-Z.M., 2006. Energy balance in phosphate calcination, paper to be published. Abouzeid, A.-Z.M., El-Jallad, I.S., Orphy, M.K., 1980. Miner. Sci. Eng. 12 (2), 7383. Abouzeid, A.-Z.M., Khazback, A.E., Hassan, S.A., 1996. In: Kemal, Arsalan, Akar, Cambazoglu (Eds.), International Mineral

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984 Processing Symposium, Izmir, Turkey, Changing Scopes in Mineral Processing, pp. 161170. Abu-Eishah, S.I., El-Jallad, I.S., Muthaker, M., Tooqan, M., Sadeddin, W., 1991. Int. J. Miner. Process. 31, 115126. Ahmed, A.A., 1990. Bull. Fac. Eng. (Assuit) Univ. 18 (2), 167170. Ahmed, A.A., 1999. 6th International Conference on Mining. Petroleum, and Metallurgy vol. 1-B, 126154. Alden, K.I., Lindgvist, I., 1964. Z. Anorg. Allg. Chem. 328, 219222. Alekseev, V.S., Shakmatova, N.U., 1964. Akad. Nauk. U.S.S.R. Kolsky Filial, Gornometallurgy Institue. Altschuler, Z.S., Cisney, E.A., Barlow, I.H., 1952. Bull. Geol. Soc. Am. 63, 12301231. Ames, L.L., 1959. Econ. Geol. 54, 829841. Ando, J., Matsono, S., 1966. Bull. Chem. Soc. Jpn. 39, 6470. Aplan, F.F., Fuerstenau, D.W., 1962. In: Fuerstenau, D.W. (Ed.), Froth Flotation, pp. 170178. AZCC, 1981. Personal Communication with Abu-Zaabal Company for Chemicals, Cairo, Egypt. Ball, D.F., Dartnell, S., Davison, J., Grieve, A., Wild, R., 1973. Agglomeration of Iron Ores, Heineman. Educational Books Limited, London, pp. 253364. Ch. 3. Bandyopadhyay, B., 1999. Preserve. Nal. USDA 61 [Preserve.Nal. usda. gov:8300/jag/v61/v61i07/610539/a610539.htm]. Bandyopadhyay, B., 2006. Personal Communication with Dr. Bhaskar Bandyopadhyay, Consultant. Baudet, G., 1988. The Processing of Phosphate Ores, Chron. Rch. Miner., Special Issue on Phosphates, p. 67. Bell, L.C., Posner, A.M., Quirk, J.P., 1973. J. Colloid and Interface Sci. 42, 124134. Bertolucci, M., 1968. In: Gould, R.R. (Ed.), Adv. Chem. Series (ACS), pp. 124135. Blazy, P., Bouhaouss, A., 2005. Miner. Metall. Process. 22 (2), 107115. Breathitt, H.W., Finch, E.P., 1977. Preprint 77-H-143, SME, AIME. Annual Meating, Atlanta. 31 pp. Bridson, D., Davies, T.W., Harrison, D.P., 1985. Clays Clay Miner. 33 (3), 258260. Cathcart, B., 1967. Proc. Escafe Seminar, Bangkok, Thailand, Minerals Resources Development Series Rep. No. 32, United nations, New York, 1968. Ciccu, R., Delfa, C., Alfanu, G.B., Carbini, P., Currelli, L., Saba, P., 1972. International Mineral Processing Congress. University of Cagliari, Italy. Coffey, R.D., Mooney, K.W., Cromwell, G.I., Aaron, D.K.-J., 1994. Animal Sci. 72, 26532660. Comonco Limited, 1974, Canadian Patent No. 939,836. Chemical and Biotechnology Abstracts in 1998. Crago, A., 1940. Process of Concentrating Phosphate Minerals, US Patent No. 2,293,640. Cuthbert, F.L., Rowland, R.A., 1947. Am. Mineral. 32, 111116. Disanto, B.J., 1985. SME Mineral Processing Handbook, Ch. 12 Pyrometallurgy, pp. 159. Ch. 12. Doss, S.K.T., 1976. J. Dent. Res. 55, 10761083. Dufour, P., Pelletier, B., Predali, J.J., Ramchin, G., 1980. Proceedings, Second International Symposium on Phosphorus Compounds, Boston, MA, p. 247. Eigeles, M.A., 1958. 4th International Mineral Processing Congress, Stockholm, pp. 391401. El-gillani, D.A., 1992. The Third Mining, Petroleum, and Metallurgical Engineering conference. Egypt, Cairo, pp. 204211. El-gillani, D.A., Abouzeid, A.-Z.M., 1993. Int. J. Miner. Process. 38, 235256. Elgillani, D.A., 1978/1979. Bulletin of Faculty of Engineering. Cairo university, pp. 237245.

83

Elgillani, D.A., Abouzeid, A.-Z.M., Negm, A.A., 1984. Final Report to the Supreme Council of Universities (FRCU) of Egypt. El-gillani, D.A., Abouzeid, A.-Z.M., 1994. 5th Int. Mineral Processing Symposium, Cappadocia, Turkey, pp. 179191. EL-Jallad, I.S., 1968. Beneficiation of the low grade hard phosphate deposits in Roseifa Area, Jordan, M. Sc. Thesis, Cairo University. El-Jallad, I.S., 1977. Investigations on the upgrading processes of the low grade phosphate. Ph. D. Thesis, Cairo University. Emich, G.D., 1984. Phosphate Rock. Ind. Miner. Rocks 2, 10171047. EPA, 1993. Phosphate Rock Processing, Mineral Products Industry, Emission Factors, Section 11. 21, pp. 110. [www.epa.gov]. Ersahan, H., Ekmekyapar, A., Sevim, F., 1994. Int. J. Miner. Process. 42, 121136. Evans, T.D., Johnston, A.E., 2004. In: Valsami-Jones, E. (Ed.), Phosphorus in environmental technologies. IWA. Faust, G.T., 1944. Econ. Geol. 39, 142151. Faust, G.T., 1950. Am. Mineral. 35, 207224. Flaszenberg, A.I., Prulov, I., Lapidot, M., 1962. Upgrading of phosphate rock. U. S. Patent 3,037,023. Freeman, H.P., Caro, J.H., Heinly, N., 1964. J. Agric. Food Chem. 12, 479486. Gu, Z., 2002. Fine Particle Flotation for Florida Dolomitic Phosphate Pebbles., M.S Thesis, College of Engineering and Mineral resources, West Virginia University. Haarr, A., 2005. The reuse of phosphorus. Eureau Position paper EU204-SL09, pp. 12. Hammoud, N.S., Khazback, A.E., Ali, M.M., 1977. International Mineral Processing Congress. Hanna, H.S., 1964. M. Sc. Thesis, Faculty of Science, Cairo University. Hanna, H.S., Somasandaran, P., 1976. In: Fuerstenau, M.C. (Ed.), A.M. Gaudin, Memorial volume, Flotation, AIME, pp. 197205. Hanna, J.S., Anazia, I., 1990. Miner. Metall. Process. 10, 8489. Hegner, P., Pacl, Z., 1974. Research Institute of Inorganic Chemistry, Czechoslovakia, Internal Report. Hendricks, S.B., Hill, W.L., 1950. Proc. Natl. Acad. Sci. 36, 731737. Henin, J.P., Lectard, A., 1983. Eng. Min. J. 10, 7784. Henin, J.P., Pinoncely, A., 1986. Ind. Miner. Min. Carr., Tech. 249252. Hignett, T.P., Doll, C.E., Livingston, O.H., Raistrick, B., 1977. Utilization of difficult ores. In: Carpentier, L.J. (Ed.), New Developments in Phosphate Fertilizer Technology, Proceedings, pp. 273288. Ince, D.E., 1987. Ph. D. Thesis, University of Florida. Jasinski, S.M., 2005. Phosphate Rock. Miner. Year book 56, 110. Jasinski, S.M., 2007. Phosphate Rock. USGS 120121. Johnston, D.J., Leja, J., 1978. Trans. IMM, pp. C237C242. Johnston, D.J., Leja, J., 1985. Trans. Inst. Min. Metall. C315C326. Khalifa, H., Orphy, M.K., Gharib, E.A., 1972. Concentration of phosphate ores, Safaga Area, MSc. Thesis., Cairo University. Kunii, D., Levenspiel, O., 1969. Fluidization Engineering. John Wiley & Sons, Inc, pp. 163. Ch. 1 & 2. Lawver, J.E., Murowchick, B.L., Snow, R.E., 1978. ISMA, Technical Economic Conferences, Orlando, Florida, Preprint TA/78/1. Lawver, J.E., Wiegel, R.L., Snow, R.E., Hwang, C.L., 1982. Min. Congr. J. 68, 27. Le Geros, R.Z., Trautz, O.R., 1967. Science 155, 14091411. Lehr, J.R., 1964. Proceedings, 17th Annual Meeting of Fertilizer Industry Round-table, Washington D.C., pp. 6167. Lehr, J.R., McClellan, G.H., 1974. Cento-Symposium on the Mining and Beneficiation of Fertilizer Minerals, Ankara, Central Treaty Organization Public relations division, pp. 194242. Lehr, J.R., Hsieh, S.S., 1981. Beneficiation of High Carbonate Ores, US Patent No. 4, 287,053.

84

A.-Z.M. Abouzeid / Int. J. Miner. Process. 85 (2008) 5984 Rosch, H., Saadi, T.A., 1970. Phosphate exploration and beneficiation studies, Jordan, United Nations, Technical Report, DP/UN/JOR70-521/2. Rudis, 1967. Mining Industrial Association TRBOVLJE, Yugoslavia. Safaga Engineering Report, vol. 2. Rule, A.R., Clark, C.W., Buttler, M.O., 1974. U.S. Bureau of Mines RI 7864. 18 pp. Rule, A.R., Kirby, D.E., Dahlin, D.C., 1978. Min. Eng. 30, 3744. Saadi, T.A.K., 1969, Mineralogy, crystal chemistry and genesis of some ores, M. Sc. Thesis, University of Durham. Saleeb, F.Z., DeBruyn, P.L., 1972. Electroanal. Chem. Interfacial Electrochem. 37, 99108. Schreider, G., H., 1981. US Patent No. 4,292,278, Sept. 1981. Sengul, H., Ozer, K., Gulaboglu, M.S., 2006. Chem. Eng. J. 122 (3), 135140. Service, A.L., Popoff, C.C., 1967. U.S. Bur. Mines Rep. RI-6935. Silverman, S.R., Fuyat, R.K., Weiser, J.D., 1952. Am. Mineral. 37, 211222. Silvia, A.F., Andery, P.A., 1972. (Jan. Feb.) Phosphorus and Potassium, vol. 57. the British Sulphur corp., London, pp. 3740. Smany, M.S., Blazy, P., Cases, J.M., 1975. Parts: I, II, III, and IV. Trans. AIME 258, 168. Smith, J.P., Lehr, J.R., 1966. J. Agric. Food Chem. 40, 342349. Snow, R.E., 1979. Beneficiation of Phosphate Ores, US Patent No. 4,144, 969. Snow, R., Zhang, P., Bogan, M., 1996. Ind. Miner. Process., Suppl. 4046. Somasandaran, P., 1968. J. Colloid Interface Sci. 27, 659665. Somasandaran, P., Agar, G.E., 1972. Trans. AIME 252, 348356. Somasandaran, P., Agar, G.E., 1976. J. Colloid Interface Sci. 24, 433440. Soto, H., Iwasaki, I., 1985. (August) Miner. Metall. Process. 160171. Steen, I., 1998, Kemira Agro, in Phosphorus and Potassium, Issue no. 217, Proceedings from JRC Workshop on problems around sludge, Stresa, 1998. Stern, S., 1965. Internnational Superphosphate Manufacturing Association, Technical Conference, vol. VI, pp. 112. Stolovitskaya, M.M., 1966. Sb. Tr. - Gos. Vses. Naucno-Issled. Inst. Stroit. Mater. Konstr. Vniistram im. P.P. Budnikova 6, 114117. Straaten, P.V., 2002. Rocks for Crops, Agro minerals of sub-Sahara Africa. (CD), ICRAF, Nairobi, Kenya. ISBN: 0-88955-512-5, pp. 724. 338 pp. Straaten, P.V., 2007. Agrogeology, The Use of Rocks for Crops. Enviroquest (pub.), (Chapter 4), pp. 87164. Taggart, A.F., 1964. Handbook of Mineral Dressing, New York. Ullmann's Encyclopedia of Industrial Chemistry, 1991, vol. A19, pp. 448455. USGS, 2003. Mineral Commodity Summaries. Jan Global Reserves and Availability. USGS, 2007. Electronic Communication with USGS Staff. Wilson, M.A., Ellis, B.G., 1984. Soil Sci. 138, 354359. Yu, Y., Xue, Y., 1996. Process for Phosphate Beneficiation, Us Patent No. 5,542,545, Aug. 1996. Zafar, Z.I., 1993. J. Nutr. Cycl. Agroecosyst. 34 (2), 173180. Zafar, I.Z., Anwar, M.M., Pritchard, D.W., 1995. Int. J. Miner. Process. 43, 123134. Zafar, Z.I., Anwar, M.M., Pritchard, D.W., 2006. Miner. Eng. 19 (14), 14591461. Zhang, P., Yu, Y., Bogan, M., 1997. Miner. Eng. 10 (9), 983994. Zhang, P., Albarelli, G.R., Stewart, K., 1999. Phosphate Beneficiation Bibliography, FIPR, Project no. 96-02-114. Zheng, X., Smith, R.W., 1997. Miner. Eng. 10 (5), 537545. Zhong, T.V., Vasudevan, A.N., Sumasandaran, P., 1991. J. Miner. Eng. 4, 563571.

Llewellyn, T.O., Davis, B.E., Sullivan, G.V., Hansen, J.P., 1982. US Bureau of Mines RI 8609. 16pp. Lovell, V.M., 1976. In: Fuerstenau, M.C. (Ed.), Flotation A.M. Gaudin Memorial volume. AIME, New York. Lu, Y., Liu, N., Wang, X., Miller, J.D., 1999. Improve phosphate flotation-nonionic polymers. In: Zhang, et al. (Ed.), Beneficiation of Phosphates, Advances in Research and Practice, Chapter 1, pp. 319. Lukomona, C., Mwalula, J.B., Witika, L.K., 2005. African Journalof Science and Technology (AJST). Sci. Eng. Ser. 6 (2), 1131129. Maslennikov, B.M., Kavitskaya, F.A., 1956. Dokl. Akad. Nauk SSSR 109, 990992. Masson, J., 1959. Enrichment by calcination of phosphate carbonate ores. Rev. Ind. Miner. 41, 651661. McClellan, G.H., Lehr, J.R., 1969. Am. Mineral. 54, 13741391. McConnell, D., Gruner, J.W., 1940. Am. Mineral. 25, 157167. Mew, M.C., 1980. World Survey of Phosphate Deposits, 4th edn. The British Sulphur Corp., London, p. 2. Miller, J.D., 2001. Improved Phosphate Flotation with Nonionic Polymers. Publication No. 02-113-150, FIPR, 40 pp. Miller, J.D., 2002. A Selective Collector for Phosphate Flotation. Publication No. 02-142-187, 60 pp. Miller, J.D., Misra, M., Yahya, A., Hu, J.S., 1987. Miner. Metall. Process. 4 (3), 133139. MIMW, 2000. Personal Communication with the Ministry of Industry and Mineral Wealth, Abu-Tartur Phosphate Project, Egypt. Mishra, R.K., 1978. Int. J. Miner. Process 5, 6976. Mishra, R.K., Chander, S., Fuerstenau, D.W., 1980. Colloids Surf. 1, 105114. Mitzmager, A., Mizrahi, J., Fischer, E., 1966. Trans. IMM C257C266. Moudgil, B.M., Chanchani, R., 1985. Metall. Process. 2 (1), 1318. Moudgil, B.M., Chanchani, R., 1986. Metall. Process. 2 (1), 1925. Moudgil, B., Gupta, D., 1989. Flotation of coarse particles. Advances in Coal and Mineral Processing Using Flotation, Engineering Foundation Conference, Palm Coast, Fl., USA, pp. 164168. Moudgil, B.M., Blanchard, F.N., Shah, D.O., Onoda, G.V., Whitney, E.D., 1986. Internal Report FIPR vol. III. Moudgil, B.M., Ince, D.E., Vasudevan, A.N., Sober, D., 1990. Miner. Metall. Process. 7 (1), 5360. Negm, A.A., Abouzeid, A.-Z.M., 2002. Utilization of East Seba'eya Phosphate Plant Taillings. Final Report to The Academy of Scientific Research and Technology. Nikitin, A.A., Vorob'ev, Yu-M., 1966. Kh. S. Sb. Tr. Gos. Vesses. Nauchn. Issled. Inst. Stroit, Materialov, Konstraktsil, No. 6, pp. 153171. Orphy, M.K., Yousef, A.A., Bibawy, T.A., 1969. Min. Mag. 121 (3), 195201. Palache, C., Berman, H., Frondel, C., 1951. Dana's System of Mineralogy, 7th ed. Wiley, New York, pp. 877889. vol. H, 1951. Perry, R.H., Chilton, C.H., Kirkpatrick, S.D., 1963. Chemical engineering handbook, GasSolid Systems, 4th edition, p. 51. Pinto, C.A.F., Yarar, B., De Araujo, A.C., 1991. SME Annual Meeting, Denver, Preprint 9180. Plotzki, E., 1937. Arch. Eisenhuttenwes. 11, 263272. Prasad, M., Majmodar, A.K., Rao, T.C., 1995. Miner. Metall. Process. 12 (2), 9296. Priestly, R.J., 1967. Proc. Intern. Symposium on Fluidization Netherlands Univ. Press, Amsterdam, p. 81. Riggs, W.F., 1989. Frothers an operators guide. In: Malhotra, Riggs (Eds.), Chemical Reagents in Mineral Processing Industry. SME, pp. 113116.

Potrebbero piacerti anche