Вы находитесь на странице: 1из 9

The reducibility of the Greek nickeliferous laterites: a review

Published by Maney Publishing (c) IOM Communications Ltd and the Australasian Institute of Mining and Metallurgy

E. Zevgolis, C. Zografidis* and I. Halikia


This paper refers to a critical review of the Greek nickeliferous laterites roasting reduction studies for better understanding of the process thermodynamic and kinetic mechanisms, affecting decisively the smelting step in pyrometallurgical extraction of nickel from these ores. From this work, it is deduced that iron and nickel oxide reduction degree does not exceed 33 and 76% respectively, the reductive reactions being stopped within the first 2030 min. Thus, part of ferric iron is transformed into ferrous iron (in the form of magnetite, wustite or fayalite) instead of metallic iron production. Also, no total conversion to metallic nickel takes place. Variation of roasting temperature (700850uC), grain size of the ores and type of solid reductants, affect the reduction rates and the final reduction degrees obtained. Diffusion or mixed control mechanisms have been found to prevail during reduction. Low reduction degrees obtained are attributed to kinetic phenomena, such as the formation of fayalite (2FeO.SiO2), which probably covers oxide grains and impedes reduction.
Keywords: Reducibility, Laterites, Review, Ferronickel, Pyrometallurgy

Introduction
Nickel is a lustrous, silvery white metal with a high melting point (1453uC), high resistance to corrosion and oxidation, good thermal and electrical conductivity, ferromagnetic properties, catalytic behaviour, ease of electroplating and excellent strength and toughness at elevated temperatures,1 properties that are indicative of the wide range of its applications. Laterite (oxide) and sulphide ores constitute the two basic natural resources of nickel, accounting for about 60 and 40% of the worlds primary nickel reserves respectively. Nickel production and demand has presented an eightfold rise since 1950 and laterite source nickel, though it formed only a small fraction of the worlds production (,10% in 1950), its participation has risen ever since to 42% (2003) and according to laterite projected economic scenarios and forecasts it is expected to rise further more to 51% by 2012.2 The three types of nickel laterite deposits with economic interest for nickel are limonitic, intermediate and saprolitic (or garnieritic). The main mineralogical phase of the saprolitic ores is garnierite (Ni,Mg)6Si4O10(OH)8 having a high magnesia and low iron content, with Ni grade 1?5 to 3?5%. Limonitic laterites contain iron oxides as the main mineralogical phase goethite (a-FeOOH,), hematite (Fe2O3) or magnetite (Fe3O4)-, thus having a high iron content

and a lower Ni grade (12%). Laterite ores, are classied into three classes:3 (i) class A: garnieritic type of laterites (Fe ,12% and MgO .25%) (ii) class B: limonitic type of laterites (high Fe content, 1532% or .32% and MgO ,10%) (iii) class C: intermediate type of laterite ores that lie between garnieritic and limonitic type of ores (Fe 1215% and MgO 2535% or 1025%).

Greek nickeliferous laterites: chemical mineralogical characterisation and industrial treatment


Greek nickeliferous laterite deposits are mainly located in three different regions: Evia Island, Lokrida in Central Greece and Kastoria in Northern Greece. Greek reserves represent 90% of nickel laterite reserves in the European Union, the remainder occurring in Finland. Evia Island and Lokrida laterites are classied in class B, while Kastoria laterite deposit is approaching a typical example of class C laterite. In Table 1, a typical chemical analysis of the Greek nickeliferous laterites is given. Evia, Kastoria and Lokrida laterite samples have been characterised by using techniques, such as XRD, differential thermal analysis and thermogravimetry, Xray uorescence, ore microscopy and electron microprobe analysis. All laterite types examined are characterised by a pissolitic texture. Evia and Lokirida laterites main mineralogical constituents were found to be hematite (Fe2O3) and quartz (SiO2), while the main nickel bearing mineral phases are the chlorite group [(Mg,Ni,Fe,Al)6(Al,Si)4O10(OH)8].4,5 Some other nickel

Laboratory of Metallurgy, School of Mining and Metallurgical Engineering, National Technical University of Athens, Heroon Polytechniou 9, 15780 Zografou, Athens, Greece *Corresponding author, email zografidis@metal.ntua.gr

2010 Institute of Materials, Minerals and Mining and The AusIMM Published by Maney on behalf of the Institute and The AusIMM Received 12 March 2009; accepted 16 August 2009 DOI 10.1179/174328509X431472

Mineral Processing and Extractive Metallurgy (Trans. Inst. Min Metall. C)

2010

VOL

119

NO

Zevgolis et al.

The reducibility of the Greek nickeliferous laterites

Published by Maney Publishing (c) IOM Communications Ltd and the Australasian Institute of Mining and Metallurgy

bearing mineral phases that have been found in selective Lokrida deposit parts are: nepouite [(Ni,Mg)3Si2O5(OH)4], and takovite [(Ni5Al4O2(OH)86H2O].5 The main mineralogical constituent of the Kastoria ore was found to be serpentine (Mg,Fe,Ni)6Si4O12(OH)6. The high goethite (aFeOOH) content of Kastoria ore indicates a higher content of the crystallic moisture, compared to the other types of Greek laterites. Serpentine and nickeliferous 2 z3 z3 magnesian crostendite (Fez 8 Fe4 )(Si4 Fe4 )O20 (OH)6 -, z2 where part of Fe can be replaced by Mg-, represent the main nickel carrier minerals in Kastoria ore.6 Microcrystallity of the nickel bearing mineral phases renders their liberation and upgrading by mineral processing techniques extremely difcult, something which would improve drastically the cost of pyrometallurgical or hydrometallurgical laterite treatment. Greek nickeliferous laterites are processed by the pyrometallurgical method Larco Process at the metallurgical plant of LARCO GMM SA, for the production of a ferronickel alloy (FeNi) with y20%Ni and low C, S and P, suitable for austenitic stainless steel production. Larco Process involves the following steps: (i) handling and mixing of raw materials (i.e. laterite, solid fuels and agglomerated rotary kiln (RK) dust in form of pellets) (ii) drying, preheating and controlled prereduction of the metallurgical mixture in RKs and production of a calcine (iii) smelting reduction of the calcine in open-bath submerged-arc electric furnaces (EFs) and production of an FeNi alloy with 1215%Ni (iv) refining and enrichment of FeNi with 12 15%Ni to FeNi with 1823%Ni in OBM converters and granulation of the final marketable alloy product.

Step 2: magnetite to wustite Fe3 O4 zCO?3FeOzCO2 Step 3: wustite to iron FeOzCO?Feo zCO2 (3) (2)

One-step reduction of nickel and cobalt oxides as follows: (a) NiOzCO?Nio zCO2 (1a) (b) CoOzCO?Coo zCO2 (1b) (i) coal de-volatilisation (ii) solid fuel (coal, lignite, coke)? carbon (C)zVolatile matter (ii) carbon gasification (Boudouard reaction) (iii) CzCO2 ?2CO (5) (4)

Mechanisms of laterite solid state reduction


The main reactions that take place during coal based reduction of iron, nickel and cobalt oxides contained in laterites, can be summarised as follows: (i) three-step reduction of hematite: Step 1: hematite to magnetite 3Fe2 O3 zCO?2Fe3 O4 zCO2 (1)

Hydrocarbons are the main gas constituents of the volatile matter. In high temperatures, the high hydrocarbons are cracked into low ones so that gaseous reductants CO and H2 are evolved.7 There is a general acceptance79 that solid-state reduction in the Fe-C-O system (in the presence of solid fuel) is mainly carried out via the gaseous intermediates (mainly CO), which are produced by the solid fuels gasication reaction (5). Regeneration of gaseous reductant through the Boudouard reaction provides the reductive atmosphere needed for the transformation of iron, nickel and cobalt oxides. Direct contact of the ore with carbon particles is probably responsible for the production of CO during the very early stages of reduction, according to the following reaction Ax Oy zC?Ax Oy{1 zCO (6)

where A5Fe, Ni or Co But the overall reductive reaction mainly occurs through the combination of reaction (7) below and reaction (5) as follows Ax Oy zCO?Ax Oy{1 zCO2 Ax Oy zC?Ax Oy{1 zCO (7)

Table 1 Typical chemical analysis of Greek nickeliferous laterites Component SiO2 Al2O3 Fe2O3 Fetot Cr2O3 MnO MgO Ni Co S CaO LOI Total Evia, % 28.2 7.0 47.5 33.2 3.1 0.04 3.2 1.03 0.05 0.4 3.0 5.0 98.88 Lokrida, % 18.6 10.9 45.0 31.4 2.7 0.04 4.0 1.15 0.06 0.45 6.6 7.5 97.41 Kastoria, % 32.2 2.9 24.8 17.3 1.4 0.01 15.4 1.45 0.06 0.40 1.45 12.5 99.03

Physico-chemical parameters affecting reducibility of nickeliferous laterites


Greek laterites
Reducibility of Greek laterites has been a subject of extended research, in order to nd optimum values of the physico-chemical parameters affecting their roasting reduction process. Given that energy requirements constitute one of the basic pillars for competitiveness and development of the metallurgical process applied, the highest necessary reduction degree (37%) of the calcine is of decisive importance, as it corresponds to important energy saving during the smelting step of the process and also because with this reduction degree,

10

Mineral Processing and Extractive Metallurgy (Trans. Inst. Min. Metall. C)

2010

VOL

119

NO

Zevgolis et al.

The reducibility of the Greek nickeliferous laterites

smelting step is a quiet process since no reduction gases are evolved through the slag in the EF. Table 2 presents a summary of the conditions employed and the main conclusions drawn by the published work on reducibility of Greek nickeliferous laterites. Temperature has proved to be one of the most important operational parameters in RK roasting reduction of the Greek laterites. According to the equilibrium diagram of the Fe-C-O system, as presented in Fig. 1, magnetite reduction (Fe3O4zCO5 3FeOzCO2) and wustite reduction (FeOzCO5 FeozCO2), can take place above 680 and 700uC respectively.10 Chemical analysis of representative samples taken along the length of industrial RKs indicated that up to 700800uC, laterite reduction remained in low levels (up to 57%) but it was drastically increased for higher temperatures11 (however, always ,30?0%). Calcine temperature is not increased above y850uC, due to agglomeration of the calcine. Moreover, all experiments carried out in the laboratory by different researchers regarding reducibility of Greek laterites with solid fuels in the temperature range of 700900uC, showed that an increase of temperature favours reduction rate of both iron and nickel oxides. Nevertheless, inuence of temperature on the reduction degree of iron oxides is positive between 700 and 750uC,12,13 or 700 and 800uC according to nickeliferous RK dust and pellets reducibility tests,14 and then it diminishes for higher temperatures. Within the same temperature range examined, nickel oxide reduction degree is increased for temperatures 700800uC and then it remains practically constant.12,13 Reducibility of laterite is considerably affected by the ore grain size, as shown during ore treatment in RKs and EFs of industrial scale, for ferronickel production.15,16 The conclusion drawn from this work is that mean size of the ore and reduction degree are inversely proportional. Indeed, even by macroscopic observation of calcined ore grains, topochemical reactions are taking place, that is, reductive reactions take place upon the ore grains surface within the temperature range 700850uC. So, by decreasing ore particle size, free surface is increased and therefore total rate of reduction is increased. It was proved that gradual decrease of feed size from 240 to 212 mm had very satisfactory results in terms of the calcines reduction degree achieved. Additionally, a ner ore corresponds to a higher mean temperature of the calcine; therefore it corresponds to a more energy saving during smelting of the calcine. Nevertheless, studying the effect of ore grain size not only in the pre-reduction but also in the smelting step of the Greek laterites, it has been concluded that a high portion of ne particles fed into the RKs favoured high levels of dust production (which means a higher nickel loss in dust, operational problems of the RKs dedusting installation and additional cost for dust handling and treatment in order to be pelletised and recycled) and was responsible for lower height of calcine-self-lining for prevention of EF walls chemical attack from slag. Laboratory studies, using Greek laterite ore and solid fuels of various grain sizes, have also conrmed that iron and nickel oxide reduction rate increases with decreasing ore and carbonaceous material grain size. This phenomenon is more evident for grain size bigger than

3 mm, as the difference in reduction obtained for ner particles is less signicant.12 The type of solid fuel used as reductant constitutes another critical parameter affecting reducibility of Greek laterites. Its role is much more complex than being just a heat source, as for example in cement production, where fuel caloric value is the ultimate critical parameter.17 In case of laterite reduction, solid fuels are utilised both for heating and reduction, thus their reactivity is of great importance. Also they can be classied according to their reactivity as follows: lignite, coal, coke and graphite.11 Lignite, is characterised by its high volatile matter, inherent moisture and low percentage of xed carbon (Cx). It evolves its thermal energy in lower temperatures compared to coal or coke and graphite, which contain a higher percentage of xed carbon and a lower percentage of volatile matter, than lignite. Laboratory studies of the Greek laterites with different types of solid reductants, maintaining the other parameters xed (temperature, ore and fuel grain size, Cx/Fe ratio, etc.) conrmed that the use of lignite18 resulted in higher iron and nickel oxide reduction degrees in comparison with the use of coal and pet coke.9 Moreover, increasing the amount of carbonaceous material relative to iron and nickel oxide content increased the rate of reduction. Also, industrial scale experiments have shown that reduction degree of Greek laterites decreases as we go from lignite to coke, in the following sequence: ligniteRcoalRcoke.11 The role of solid fuels used for reduction of Greek laterites in RKs has also been studied.11,16 The conclusions drawn after extensive evaluation of operational data is that optimum results with respect to calcine reduction are obtained in the appropriate combination of solid fuels, so that the necessary amount of volatiles and Cx respectively is contained. Moreover, increase of the lignite amount has proved to result in an increased reduction rate, due to burning of the volatile matter in the rst zone of the RKs (drying, preheating zone) and therefore maintaining of the required reductive atmosphere, so that a higher reduction is obtained. Nevertheless, very high content of volatile matter in the metallurgical mixture corresponds to high temperature of RK ue gases, which cause increased thermal losses and risk for re in the dedusting installation. For this reason, decrease of lignite participation in the metallurgical mixture is required and increased amount of less reactive coal with higher content of the main reductive agent Cx, resulting in a lower specic consumption of fuel per ton of natural laterite. There are various approaches published in international literature concerning the effect of solid fuel volatile matter, in iron ore reduction. All of them however seem to agree that volatiles increase the rate of reduction. There is also an agreement among different researchers that participation of volatiles in reduction depends mainly on temperature. Thus, at temperatures below 9001000uC,7,8 reduction by volatile matter is regarded as quite signicant, whereas at higher temperatures the role of volatiles is negligible, so that at higher temperatures reduction mainly takes place through solid carbon. This is due to the fact that most of volatiles are evolved at temperatures ,1000uC.11 Liu et al. and Strezov et al.,19,20 also investigated the fundamental mechanisms for iron ore nes reduction

Published by Maney Publishing (c) IOM Communications Ltd and the Australasian Institute of Mining and Metallurgy

Mineral Processing and Extractive Metallurgy (Trans. Inst. Min. Metall. C)

2010

VOL

119

NO

11

Published by Maney Publishing (c) IOM Communications Ltd and the Australasian Institute of Mining and Metallurgy
Zevgolis et al.

12

Table 2 Laboratory and industrial results on reducibility of Greek nickeliferous laterites Halikia and Skartados18 EviaLokridaKastoria: 602515 wt-% 26z0.038 mm 2208z37 mm EviaLokridaKastoria: 602515 wt-% Kastoria Neou et al.9 Neou et al.13 Zevgolis et al.14 Greek Ni-ferrous laterite rotary kiln dust Dust: 2105 mm; pellets: 9 mm

Reference

Zevgolis11

Halikia-Neou et al.12

Origin of laterite ore

EviaLokridaKastoria: 602515 wt-%

The reducibility of the Greek nickeliferous laterites

Ore grain size

EviaLokridaPellets (from rotary kiln dust): 70305 wt-% Ore: 215 mm; pellets: 212 mm Lignite, coal 1 : 6.021 : 5.01 26z0.038 mm Pet coke 1 : 4.63 2208z37 mm Ore: 20.250z0.038 mm; pellets: 210z5.6 mm, 214z10 mm Coal 1 : 3.51 20.150z0.106 mm

Reductant Cfix/Fetot Reductant grain size Bulk ore and solid fuels 860900 29.55% 900uC Lignite 750900 23.16% 900uC Ground laterite and pet coke

Coal, lignite, Cfix (of dust) 1 : 5.24 230 mm, 25 mm, 2100 mm

Cfix (of dust) 1 : 2.41 : 2.6

Mineral Processing and Extractive Metallurgy (Trans. Inst. Min. Metall. C)

Specimen form

Bulk ore and solid fuels

2208z38 mm, 2100z38 mm, 253z38 mm Lignite 1 : 4.31 2208z147 mm, 2147z104 mm, 2104z74 mm Ground laterite and lignite

Rotary kiln dust pellets (of same dust) 700850 28% 850uC Rotary kiln dust

2010

Temperature, uC Maximum reduction degree conditions

920 25% 920uC 215 mm ore, 212 mm pellets, 230 mm coal, 25 mm lignite

700850 32.73% 850uC 253z38 mm ore 2104z74 mm lignite

Pulverised ore pellets binding agent: bentonite, 1% 700900 31.74% 900uC Pulverised ore

VOL

119

72.4 (iron oxide) Diffusion (iron oxide) Mixed kinetic model: chemical reaction diffusion (nickel oxide)

NO

Activation energy E, kJ mol21 Rate controlling step

40.17 (iron oxide) 87.45 (nickel oxide) Diffusion (iron oxide) Mixed kinetic model: chemical reaction diffusion (nickel oxide)

42,E,84 (700750uC) E,42 (800850uC) Mixed kinetic model: chemical reaction diffusion (iron oxide)

Zevgolis et al.

The reducibility of the Greek nickeliferous laterites

Published by Maney Publishing (c) IOM Communications Ltd and the Australasian Institute of Mining and Metallurgy

1 Equilibrium curves of CO-CO2 versus temperature for Fe-C-O system10

in coalore mixtures and the role of the fuel volatile matter. They reported that volatile matter constituents were methane (CH4) within the temperature range 450 700uC and CO kai H2 for temperatures .800uC. Gas agents CO kai H2 are mainly responsible for hematite (Fe2O3) reduction up to 800uC, while additional CO generated due to the Boudouard reaction at elevated temperatures resulted in acceleration of the reduction procedure. Concerning laboratory work with Greek laterite, the fact that remaining carbon in the calcine is in most cases much higher than the theoretically expected, indicates according to researchers that evolved volatiles participate in the oxide reduction,9,12 but this requires further investigation. Reducibility tests have been conducted by different researchers, on Greek laterites in the form of dust and pellets as well. Dust produced during industrial treatment of laterite in the RKs has a higher Ni content than the ore and a carbon content .8%, which is higher than the stoichiometrically required (y6?5%) for FeNi 13 production. Thus, it is evident that agglomeration and recycling of dust, which is y7% of the laterite feed, contribute to the economics of the metallurgical process. So, the reductive behaviour of the industrially produced cement bonded laterite pellets has been studied for comparison with the behaviour of same origin laterite

dust.14 Reducibility tests of pulverised Kastoria laterite ore have been also conducted using coal as a reductant, in order to compare with the reductive behaviour of agglomerated mixture (pellets) of Kastoria pulverised ore and coal (the stoichiometrically required for FeNi 13 production) with bentonite as the binding agent.13 The temperature range of the experimental procedure was 700900uC. It was deduced from both series of tests, that iron oxide reduction degree obtained for pellets was lower than for dust (with same laterite) under the same conditions, though no serious difference is observed concerning nickel oxide reduction. This difference in the reductive behaviour between laterite dust and pellets can be attributed to the fact that the binding agent covers part of the oxide grains decreasing their reactive surface area. The conclusion drawn from all reducibility studies of Greek nickeliferous laterites, either conducted in an industrial or in a laboratory scale, within the temperature range of 700900uC, is that independently of the conditions employed, i.e. ore grain size, temperature and type and amount of the solid reductant, in no case calcine reduction degree exceeds 33%, though the remaining carbon in the calcine is almost always higher than the theoretically required for further progress of reductive reactions. Given that the conversion of Fe2O3

2 Reduction degree versus time typical diagrams at various temperatures of a laterite ore mixture (Evia IslandLokrida Kastoria: 602515 wt-%, 20?053z0?037 mm) with lignite (20?149z0?0940 mm),12 b electrostatic lter dust14 and c electrostatic lter pellets14

Mineral Processing and Extractive Metallurgy (Trans. Inst. Min. Metall. C)

2010

VOL

119

NO

13

Zevgolis et al.

The reducibility of the Greek nickeliferous laterites

Published by Maney Publishing (c) IOM Communications Ltd and the Australasian Institute of Mining and Metallurgy

3 Temperature and reduction degree in RK

to Fe3O4 and to FeO corresponds to a reduction of 11?1 and 33?3% respectively, it comes that magnetite and wustite should coexist in the calcine and no further conversion to metallic iron exists. Additionally, it is concluded that reductive reactions occur within the rst 20 to 30 min of the process and then they practically stop, for all cases examined. This can be attributed to kinetic phenomena, such as the formation within the temperature range examined of iron-silicate minerals, such as fayalite (2FeO.SiO2), which probably covers oxide grains, thus impeding progress of the reduction. In Fig. 2, typical reduction degree diagrams versus time within a temperature range 700850uC are presented, for reduction of pulverised laterite ore mixture (Evia islandLokrida Kastoria: 602515 wt-%, 20?053z0?037 mm) with lignite (20?149z0?0940 mm), and also the reduction of electrostatic lter dust from RKs as well as the reduction of laboratory made pellets from this dust. From these diagrams it comes that reduction degree does not exceed 33%. The same is shown in Fig. 3, with reduction in industrial  RKs. (In Fig. 3, reduction is dened as the ratio Fe2z Fetot |100, so the real reduction degree, i.e. Oi { Of =Oi |100, where Oi is the initial oxygen and Of is the nal oxygen in iron oxides, is ,33?3%.) Industrial operation with garnierite type of ores shows that reduction in the RK goes up to 37%, which means that we have metallic iron in the calcine.21

World experience
Reducibility tests concerning different types of nickeliferous laterites according to international literature,22,23 indicate that chemical mineralogical composition of the

laterite ore is a critical parameter for the nal degree of reduction obtained. Reducibility of several types of nickeliferous laterites from the Dominican Republic22 (grain size, 20?074 mm) has been studied within the temperature range 4001000uC, using hydrogen as a reductant. The aim of the study was to compare the reductive behaviour of garnieritic, limonitic and intermediate type of ores. The experimental results clearly indicated that reducibility of each oxide (iron, nickel and cobalt oxide) depends on the ore type. Moreover, chemical and mineralogical analysis of the reduced samples indicated that nickel and iron (oxide) degree of metallisation was higher in limonitic than in garnieritic and intermediate type of ores. Specically, both iron and nickel oxide degree of metallisation obtained after 40 min processing of the garnieritic type of ore at 1000uC was y10%. Temperature variation between 400 and 1000uC did not affect signicantly the reduction process. The respective degrees of metallisation were 80 and 70% at 1000uC concerning limonitic type of ore and they were dramatically increased .800uC. Cobalt degree of metallisation at 1000uC was slightly lower (y65%) in limonitic compared to garnieritic type of ore (70%) and the effect of temperature variation was negligible concerning the rst and signicant .850uC concerning the second one. This is different from what happens in reduction by solid fuels. That is, in solid fuel reduction limonitic type of ores cannot be reduced to metallic iron, so reduction stops when reduction degree of iron oxides is ,33?3%. Low nickel oxide reducibility in the garnieritic type ores, was attributed to the formation of olivine, a nickeliron magnesium orthosilicate and the tendency of nickel to

14

Mineral Processing and Extractive Metallurgy (Trans. Inst. Min. Metall. C)

2010

VOL

119

NO

Zevgolis et al.

The reducibility of the Greek nickeliferous laterites

Published by Maney Publishing (c) IOM Communications Ltd and the Australasian Institute of Mining and Metallurgy

exchange places with magnesium in silicates, which are stable in high temperatures. On the contrary, the low magnesia and silica content of the limonitic type of ores is not adequate to result in hosting nickel in the olivine phase, thus the per cent of nickel oxide reduction degree increases for high temperatures. Cobalt oxide reducibility was higher for the garnieritic laterite type and the effect of temperature was more evident for lower temperatures, due to the cobalt tendency to replace iron in the limonite lattice. Reductive reactions concerning all types of laterite ores examined, after 40 min practically stopped. In the Greek laterite reducibility tests, as it has already been mentioned above, reductive reactions practically stopped after 20 to 30 min. Reducibility tests of a garnieritic type of laterite ore in the form of pellets, in the range 7001000uC, using a CO/ CO2 mixture as a reducing agent has also been conducted.23 Olivine (Mg,Fe)2SiO4 formation was proved to be critical for the reduction progress, since it is stated that reducibility (determined by percentage of weight loss after rst calcined in the same temperature laterite ore pellets) reaches the highest possible values for temperatures above serpentine (Mg, Fe, Ni)6Si4O12(OH)6 decomposition (y600uC) and below transformation temperature of amorphous olivine to a stable mineralogical phase (y810uC). It was also deduced that a strongly competitive relation exists among reduction progress and olivine formation, i.e. a slow reduction rate (by employment of mild reducing atmosphere through gas reducing agent) relative to olivine formation results to a lower reduction degree. On the contrary, rapid reduction rate (by employment of intensive reducing atmosphere during the rst minutes of the reductive procedure) relative to the crystallisation of the olivine phase results in higher reduction degree values.

Linearity assessment of diagrams ln[2ln(12a)] versus lnt and determination of the slope n, were obtained through the application of the least squares method. It is noted that no time greater than 15 min was used for the kinetic analysis, given that after the rst 20 min the reactions tended to stop, in all cases examined. The conclusion deduced from this work was that diffusion kinetic equations (D1)(D5) best tted the experimental data concerning iron oxide reduction. Moreover, it has been reported that the ValensiCarter equation was the mathematical model that best represented the proposed diffusion mechanism, with Z value equal to 0?5. Another value of coefcient Z would probably result in an even more representative mathematical model. With respect to the nickel oxide reduction kinetics, it was reported that chemical reaction is the rate controlling step for the lower temperatures (up to 800uC) but no certain kinetic model from Table 3 (equations (F1), (R1) and (R2)) can be stated to be the best tting, due to the fact that all of the aforementioned three equations t very well. The rate controlling step though, changes for higher temperatures (up to 900uC), but there has not been a conclusion based on reported approaches about the mechanism that prevails within the temperature range 800900uC. Reduction kinetics of RK dust and laboratory made pellets of the same origin, in the temperature range 700 850uC, were studied.14 The experimental data obtained concerning iron oxide reduction, were applied to the following mathematical models: (i) CrankGinstlingBrounshtein (CGB) kinetic model 1{(2=3)a{(1{a)2=3 ~Kt (:D4) (9)

rate controlling step: diffusion through the product layer (ii) {(1{a)1=3 ~Kt (:R2) (10) rate controlling step: chemical reaction at the interface between the unreacted core and the product layer 1{(2=3)a{(1{a)2=3 z1{(1{a)1=3 (11) ~Kt (:D4zR2) Generalised equation, that is a combination of equations (9) and (10), based on the additivity of reaction times. rate controlling step: mixed controlled mechanism.
Table 3 n values reactions Kinetic equation of kinetic equations n for gassolid

Reduction kinetics of Greek nickeliferous laterites


Kinetic analysis carried out for the Greek nickeliferous laterite roasting reduction, is based on the unreacted shrinking core model. Roasting reduction kinetics of Greek laterite ne particles (EviaLokridaKastoria ore mixture: 602515 wt-%, granulometry 20?250z 0?037 mm, temperature range 700900uC) with lignite and pet coke as reductants respectively, were conducted.9,12 The methodology of work used for approaching the rate controlling step of the process, is the application of the diagnostic equation ln{ln1{a~n ln tzln b (8)

(iii)

where a is reduction degree (%) of iron or nickel oxides, t is time (s), b is constant and n is constant depending on the rate controlling mechanism and the geometrical characteristics of the ore and solid reductant particles. The obtained n values from application of the experimental data at, which represent the slopes of the linear graphic representation of equation (8), are compared with the theoretical values of the widespread used kinetic equations of Table 3. It is noted that equations (D1) (D5) correspond to the diffusion rate controlling step, equations (F1), (R1) and (R2) correspond to chemical reaction mechanism and equations (A2) and (A3) correspond to the nucleation rate controlling step.

0.62 D1: a25Kt D2: (12a)ln(12a)za5Kt 0.57 0.54 D3: [12(12a)1/3]25Kt 0.57 D4: 12(2/3)a2(12a)2/35Kt D5: Kt5{Z*z[1z(Z21)a]2/32(Z21)(12a)2/3}/(Z21) (Valensi Carter equation) F1: 2ln(12a)5Kt 1.0 1.11 R1: 12(12a)1/25Kt 1.07 R2: 12(12a)1/35Kt 2.0 A2: [2ln(12a)]1/25Kt 2.0 A3: [2ln(12a)]1/35Kt *Z: coefficient representing the product volume per volume of the reactants consumed. Z is assumed to be 0?5.

Mineral Processing and Extractive Metallurgy (Trans. Inst. Min. Metall. C)

2010

VOL

119

NO

15

Zevgolis et al.

The reducibility of the Greek nickeliferous laterites

Published by Maney Publishing (c) IOM Communications Ltd and the Australasian Institute of Mining and Metallurgy

It was deduced from the kinetic analysis that the mixed kinetic model equation (11), ts well the experimental data for all dust samples and pellets of a certain origin (the washing tower) up to 750uC. This kinetic model also veried the experimental data for pellets of different origin (i.e. electrolter and polycyclone dusts) up to 800uC. At higher temperatures, i.e. up to 850uC, it has been conrmed the predominance of diffusion mechanism, according to the CGB equation (9). The activation energy value of the rate-determining step at various conditions has been evaluated, concerning coal based laterite reduction, by means of the widespread rate expression (12) that follows r~kC n (12)

either on its own or in combination with chemical reaction in the interface product layer gas reactant. Kinetic analysis of Greek nickeliferous laterites enhances the conclusion drawn by all the reported reducibility tests, according to which iron silicate mineralogical phases formed in the temperature range examined (700900uC), like fayalite (2FeO.SiO2), probably cover the iron oxide grains and thus impede the progress of the reaction.

Conclusions
The main physicochemical parameters affecting reducibility of the Greek nickeliferous laterite ores have been reviewed. It is concluded that no matter if controlling the values of physicochemical parameters, such as temperature, grain size of materials and type of solid reductant to be optimum, iron and nickel oxide reduction degrees obtained do not exceed 33 and 76% respectively, though the remaining carbon in the calcine is adequate for further progress of the reductive reactions. This means that practically no metallic iron is formed during roasting reduction of Greek laterites and nickel oxide is partially transformed to metallic nickel. Decrease of the ore and solid reductant grain size (mainly for granulometry z3 mm) and use of reactive solid fuels (lignite) instead of less reactive (coal or coke), favour considerably iron and nickel oxide degree of reduction. Increase of temperature within the temperature range 700900uC examined, results in increase of oxides reduction rate, but its inuence on iron and nickel oxide degree is evident up to 750800uC and then it diminishes. Reduction process progresses initially with a relatively high rate for the rst ten minutes, then the rate starts to decrease, until it diminishes to zero after the rst 20 min reaction. The conclusion that reductive reactions practically stop after a time of 2040 min is veried by published approaches concerning reducibility studies of different origin laterite ores. It can be attributed to the formation within the temperature range examined of iron silicate minerals, such as fayalite (2FeO.SiO2) or forsterite (Mg2SiO4), which probably cover oxide grains and impede further progress of the reduction. Thus, future reducibility studies of Greek nickeliferous laterites should be focused on mineralogical analysis of the reduction products in relation with the factors affecting the formation and reduction of complex nickelironmagnesium silicate phases, so that the highest possible reduction degree is obtained. Kinetic studies of the reductive procedure showed that mixed control and diffusion mechanisms have been found to prevail during reduction of oxides.

where r is the reaction rate, k the rate constant, C the uid reactant concentration (mol L21) and n the order of the reaction. Assuming that the uid reactant concentration is constant, the Arrhenius law [lnk5lnAoE/R(1/T)] can be used for calculation of the activation energy through slope determination of the following linear equation lnr~ln(Ao C n ){(E =R)(1=T ) (13)

where r is the mean initial reduction rate value, calculated through the expression r5DaDt21 (a is the reduction rates obtained from the experimental data), Ao is the frequency factor of the Arrhenius law, C is the uid reactant concentration (mol L21), n the order of the reaction, E the activation energy, T the temperatures (K) examined for the kinetic analysis and R the universal gas constant. Activation energy has also been calculated by Zevgolis et al., concerning RK dust and laboratory made pellets of the same origin, through application of the Arrhenius law to the slopes of the generalised equation (13) lines, mixed controlled mechanism, at 700 and 750uC, i.e. by means of the relationship   ln k(T2 ) E 1 1 (14) ~{ { ln k(T1 ) R T2 T1 The activation energy values concerning iron oxide reduction9,12 (72?4 and 40?2 kJ mol21), indicate that the prevailing kinetic mechanism may be either the mixed controlled mechanism (diffusionchemical reaction) or diffusion. Activation energy value reported concerning nickel oxide reduction (87?4 kJ mol21), indicates the predominance at least for a certain temperature range of chemical reaction mechanism. The mixed controlled mechanism for iron oxide reduction was proposed by Zevgolis et al. for low temperatures (700750uC) and diffusion through the product layer for higher temperatures (800850uC).14 The conclusions drawn from the kinetic analysis of the Greek nickeliferous laterite solid state reduction, are in agreement with the conclusions drawn by other studies, concerning reducibility tests of different origin laterite samples, either in pellet form23 or in form of cement bonded laterite briquettes with CO-CO2 gas reducing mixture,24 within the temperature range (700 1000uC). Mixed control (chemical reaction and diffusion of the reducing gas agent through the product layer), is the mechanism that seems to prevail during the reductive procedure. It is apparent from the aforementioned, that diffusion constitutes a serious kinetic factor affecting solid state reduction of laterites,

References
1. www.lib.murdoch.edu.au/adt/pubfiles/adt-MU20051004?114504/ 02/Whole.pdf 2. A. D. Dalvi, W. Bacon and C. Osborne: The past and the future of nickel laterites, Proc. PDAC 2004 Int. Convent., Toronto, Canada, March 2004, PDAC, 127. 3. E. N. Zevgolis: Extractive metallurgy of nickel: part I. Pyrometallurgical methods; 2000, Athens, National Technical Univerity of Athens. 4. E. Mposkos, A. Orfanoudaki and Th. Perraki: The Ni distribution in the mineral phases of Greek FeNi laterite deposits, Proc. 3rd Symp. on Mineral wealth, Athens, Greece, November 2000, Technical Chamber of Greece, 107115. 5. N. Albadakis: Ni-minerals in the deposits of the sub-pelagonic zone, Miner. Wealth, 1984, 31, 932.

16

Mineral Processing and Extractive Metallurgy (Trans. Inst. Min. Metall. C)

2010

VOL

119

NO

Zevgolis et al.

The reducibility of the Greek nickeliferous laterites

Published by Maney Publishing (c) IOM Communications Ltd and the Australasian Institute of Mining and Metallurgy

6. S. Agatzini: A new approach to the metallurgical treatment of nickeliferous laterites, Report within the framework of the CEC BRITE-EURAM Programme ECU 368?000 (In cooperation with University of Hertfordshire, University of Minho), 1993, 1 12. 7. Q. Wang, Z. Yang, J. Tian, W. Li and J. Sun: Mechanisms of reduction in iron orecoal composite pellet, Ironmaking Steelmaking, 1997, 24, (6), 457460. 8. E. Donskoi, D. L. S. McElwain and L. J. Wibberley: Estimation and modeling of parameters for direct reduction in iron ore/coal composites: part II. Kinetic parameters, Metall. Mater. Trans. B, 2003, 34B, 255266. 9. P. Neou-Syngouna, I. Halikia and K. Skartados: Prereduction of laterites with petroleum coke: influence of the granulometric size on its progress and kinetics, Min. Metall. Ann., 1997, 1, 25 49. 10. E. N. Zevgolis: Iron-cast iron metallurgy theory and technology; 2004, Athens, National Technical Univerity of Athens. 11. E. N. Zevgolis: A contribution to the study of problems of rotary kilns in roasting reduction of greek nickeliferous laterites, Thesis for lectureship, National Technical Univerity of Athens, Athens, Greece, July 1982. 12. I. Halikia, P. Neou-Syngouna and M. Katapotis: Reductive roasting of iron-nickel ore using greek lignite thermodynamic and kinetic approach, In honor of Professor Emeritus of NTUA A. Z. Fragiskos; 1998, Athens, National Technical Univerity of Athens. 13. P. Neou-Syngouna, I. Halikia, C. Skartados, L. Papadopoulou and G. Portokaloglou: Comparative study of laterite roasts in the form of powder and pellets, Min. Metall. Ann., 1999, 12, 85 118.

14. E. N. Zevgolis, I. Halikia and I.-P. Kostika: Reductive behavior of the recycled dust during nickeliferous laterite treatment, Erzmetall the World of Metallurgy, 2006, 59, (6), 350359. 15. E. N. Zevgolis: The importance of iron ore grain size in rotary kiln operation, Miner. Wealth, 1986, 45, 103110. 16. E. N. Zevgolis: The ore grain size effect in ferroalloys production by the rotary kiln electric furnace method, Miner. Wealth, 1988, 54, 3946. 17. E. N. Zevgolis and A. Tzamtzis: The role of solid fuels used for reduction in rotary kilns, Techn. Chron. C, 1987, 7C, (2), 519. 18. I. Halikia and K. Skartados: Effect of solid reductant on the metallurgical behaviour of laterite calcine, In Memory of NTUA Professor Emeritus J. Papageorgarakis, 294303; 2001, Athens, National Technical Univerity of Athens. 19. G.-S. Liu, V. Strezov, L. A. Lucas and L. J. Wibberley: Termal investigations of direct iron ore reduction with coal, Thermochim. Acta, 2004, 410, 133140. 20. V. Strezov, G.-S. Liu, J. A. Lucas and L. J. Wibberley: Calorimetric study of the iron ore reduction reactions in mixtures with coal, Ind. Eng. Chem. Res., 2005, 44, 621626. 21. E. N. Zevgolis: Report from visit to FENIMAK Ferronickel Plant, Kavadarci, Fyrom, 24 February 2000. 22. M. Kawahara, J. M. Toguri and R. A. Bergman: Reducibility of laterite ores, Metall. Trans. B, 1988, 19B, 181185 23. S. Li and K. S. Coley: Kinetics and mechanism of reduction of laterite ore high in serpentine, Proc. J. M. Toguri Symp. on Fundamentals of metallurgical processing, (ed. G. Kaiura et al.), 179192; 2000, Ottawa, CIM. 24. H. Purwanto, T. Shimada, R. Takahashi and J. Yagi: Reduction rate of cement bonded laterite briquette with CO-CO2 gas, ISIJ Int., 2001, 41, S31S35.

Mineral Processing and Extractive Metallurgy (Trans. Inst. Min. Metall. C)

2010

VOL

119

NO

17

Вам также может понравиться