Sunteți pe pagina 1din 16

XRD analysis and evaluation

ofiron ores and sinters 3


J.P.R. de Villiers1, L. Lu2
1
University of Pretoria, Pretoria, South Africa
2
CSIRO Mineral Resources Flagship, Pullenvale, QLD, Australia

3.1Introduction
The iron and steel industry is confronted by decreasing grades and increased impurity
levels and a buildup of lower-grade discard material that needs to be upgraded. In
modern blast furnaces, stronger, more reactive sinters are required for a high produc-
tivity. To address these problems, knowledge of the phases and their impurity levels is
necessary so that the most appropriate measures can be devised.
X-ray powder diffraction (XRD) has a unique capability to distinguish between
different phases on the basis of their crystal structure and their chemistry. This has
been extensively utilized following the development of suitable instrumentation and
has led to the characterization and examination of powdered or fine-grained materials
for their phase composition. The compilation of the data into a database of known
diffraction patterns has led to widespread acceptance of the X-ray powder diffraction
method for phase identification. A major limitation however was the lack of quantita-
tive capability.
With the development of fast detectors, modern accurate instrumentation, and ad-
vanced software, X-ray diffraction analysis has undergone a revival. It is used increas-
ingly in the field of mineral processing to characterize minerals, to track their behavior,
and to optimize the performance of metallurgical processes. It is fast, requiring no spe-
cial sample preparation, apart from fine grinding, and together with statistical methods
such as cluster analysis, can provide valuable information to the process engineer. It
is usually the first method, together with X-ray fluorescence and microanalysis, to be
used for the characterization of iron ores and products arising from their processing.
SEM methods used for mineral quantification are not well suited for the distinction of
hematite and magnetite in ores and especially in sinters, so again, XRD is the method
of choice for phase quantification.
All mineral processing operations treat minerals (not elements), and as such, their
identity, their properties, and their behavior in these processes are of importance. The
minerals can also vary spatially across an ore body, and it is therefore necessary to
characterize the ores and to measure their distribution in order to optimize mining and
mineral processing operations.
In high-temperature processes, the behavior of the feedstock is of importance in the
calculation of phase equilibria, slag formation, and mass and energy balances. In all

Iron Ore. http://dx.doi.org/10.1016/B978-1-78242-156-6.00003-4


2015 Elsevier Ltd. All rights reserved.
86 Iron Ore

these calculations, the phase compositions and their quantities are important variables
to be used in thermodynamic modeling of the processes. In iron ore sinters and pellets,
the identity of the bonding phase is of importance, and this influences the strength and
reducibility of the feedstock in the subsequent blast furnace operation.
In this contribution, after a short theoretical analysis, the applications of the method
will be discussed, followed by practical applications in iron ore processing.

3.2Principles of powder X-ray diffraction


Powder X-ray diffraction (XRD) analysis uses a monochromatic beam of X-rays to irra-
diate a sample with the X-ray source and detector making variable angles with normally
a flat sample. The diffracted X-rays from the crystal planes in the powdered phases that
will be measured by a detector will show up as peaks of variable intensities at specific
diffraction angles. Each phase or mineral will have a unique diffraction pattern that is a
function of its crystal structure, its crystallite size, and its lattice strain, if present.
The intensity Ihkl of a diffraction peak is a function of a number of factors and is
expressed in abbreviated form as the following equation:

2
I hkl = S M hkl LP A Phkl Fhkl

where S is the scale factor that equates the experimental diffraction pattern to the cal-
culated pattern. It is a function of the crystal structure of the phase and the quantity
of the phase in a mixture, Mhkl is a multiplicity factor that depends on the number of
times that symmetrically equivalent reflections overlap in a diffraction pattern, LP is
a combined term depending on the geometry of diffraction and the partial polarization
of the X-ray beam, A is a correction for the absorption of the X-ray beam, Phkl is the
preferred orientation correction term that corrects for possible deviations from the
random orientation of crystallites in the sample due to their shape, Fhkl is the structure
factor that is defined by the crystal structure.
The structure factor is a function of the scattering power of the constituent atoms
and their positions and is given by
N
2p i ( hx + ky + lz )
Fhkl = f n e
n =1

where fn is the atomic scattering factor of the nth atom; x,y,andz is the position of the
nth atom.
The temperature factor B, also called the DebyeWaller factor, is an additional factor
2
sin q
-B
l
affecting the intensity so that f n = f 0 e
for each atom position for each phase.
The diffraction peaks are further affected by instrumental and sample effects that
cause broadening so that the combination of all these gives the peak profiles that are
observed in the experimental diffraction pattern.
XRD analysis and evaluation ofiron ores and sinters87

This discussion on the theory of X-ray diffraction analysis is kept to a mini-


mum, since this can be sourced from a number of excellent references by Klug and
Alexander (1974), Jenkins and Snyder (1996), Pecharsky and Zavalij (2005), and ref-
erences therein. The latter reference is a very useful modern treatise on the elements of
powder diffraction. The real purpose of this contribution is to describe the versatility
of the method and its application in iron ore processing.
The diffraction patterns of a very large number of phases and minerals are tabu-
lated in a database (Powder Diffraction File, issued by the International Centre for
Diffraction Data (ICDD)) consisting mainly of interplanar d-spacings and intensi-
ties of the diffracting planes. The diffraction pattern of the phase being investigated
can then be compared with those of the phases in the database to identify it and
to get an approximation of its quantity in a mixture of phases. Figure3.1 shows a
schematic diagram of the most common geometry used in modern XRD equipment.
The X-ray beam from the X-ray source is directed onto a flat sample, while the
source and detector scan increasing angles towards the vertical. A set of slits ensure
that the beams are focused. In modern instruments, a solid-state strip detector is
used that increases measurement speed almost a hundredfold compared with the
older point detectors.
The output consists of a list of angles and their intensities collected from the sample
being investigated. A typical XRD pattern of a goethitic iron ore is shown in Figure 3.2.
The positions of the peaks calculated from crystal structure data are shown as vertical
lines. For example, the mineral content of a goethite containing iron ore can readily
be identified and quantified by XRD with a 10min scan. Much additional information
relating to the element distribution in the various minerals can be obtained when using
powder XRD together with mineral compositions obtained by SEM analysis.

Detector

Mono-
chromator
Divergence slit Antiscatter
slit

Detector
X-ray slit
source
Sample

Figure3.1 The BraggBrentano geometry used in modern diffractometers.


Courtesy of Bruker AXS.
88 Iron Ore

Counts
DeVilliersJ_FCF11A

15,000

10,000

5000

0
10 20 30 40 50 60 70 80
Position (2q ) (Cobalt (Co))

Peak list

010791741; Hematite, syn; Fe2 O3

030650466; Quartz low, syn; SiO2

010810464; Goethite, syn; FeO(OH)

010861342; Magnetite; Fe2.886 O4

Figure3.2 XRD pattern of an iron ore sample, showing the peak positions of the major
minerals. Extensive peak overlap is apparent.

3.3Rietveld analysis
With the advent of modern fast computers, the fitting of the experimental XRD pattern
with that of the calculated one has become possible, and the method is being used for
the structure refinement of unknown phases and for the quantification of phases or
minerals in a mixture. The experimental XRD pattern is compared with the pattern
calculated from the crystal structures of the constituent phases, and a measure of the
goodness-of-fit and difference curve are calculated. This is done for every analysis
step in the diffraction pattern. The least squares method then calculates a number of
scale factors that are directly proportional to the relative amounts of the phases. A
number of discrepancy factors are also calculated. A useful treatise of the method is
given in the monograph edited by Young (1993). Discussions on phase quantification
using the method are given in numerous articles, for example, by Madsen etal. (2001).
Necessary inputs for Rietveld analysis are the following:
An aligned X-ray diffractometer, calibrated with an internal standard
An experimental XRD pattern, measured in steps of at least 0.02 degrees 2
Accurate cell dimensions, space group, and atomic positions for each phase
A sample, micronized to a particle size of approximately 1015m
Sample rotation, essential to lessen the effects of preferred orientation
Suitable software for the analysis

Rietveld software programs are designed to describe the peak profiles as accurate as
possible. The XRD peaks can either be fitted to some analytic function d escribing the
peaks (usually some variant of a Gaussian or Lorentzian distribution that has limited
XRD analysis and evaluation ofiron ores and sinters89

W
Y(2q ) = (W G) S
0 +
Emission profile
tan(q )

G
0 + 0 + 0+ 0+
Target Slit width Horizontal Axial
divergence divergence
a 2 cot(q ) SL2 cot(q ) 0 +
Final profile
Y(2q )

S
0+ 0 +
Absorption Crystallite size Strain
sin(2q )/AB 1/cos(2q ) CS STR tan(q )

Figure3.3 Convolution of the various parameters that influence the peak shapes showing
their contribution to the final peak shape.
Courtesy of Bruker AXS.

physical meaning) or be calculated using the so-called fundamental parameters.


Normally, for software using the fundamental parameters approach, the instrumental
parameters for the diffractometer must be determined using a suitable standard. This is
shown pictorially in Figure3.3. As can be seen, the diffraction peaks are based on the
convolution (folding) of the X-ray emission profile, W, the instrumental parameters,
G, and the sample parameters, S. The X-ray emission profile is a given for a specific
radiation and is taken from previous determinations. The instrumental parameters are
determined once for a specific instrument using a well-characterized XRD standard
such as LaB6. Once these are fixed, then only the sample characteristics such as the
crystallite sizes, the lattice strain, and the relative amounts of the phases are refined.
So, in Figure3.3, the parameters W and G are essentially fixed, leaving only the sam-
ple parameters S to be refined.
For quantitative analysis, Hill and Howard (1987) had derived simple relationships
on which the Rietveld method relies for quantification without or with an internal
standard. These are given as
S P ( ZMV ) P
WP = without an internal standard,
( )
n
i =1
S i ZMV i assuming no amorphous phase present,
and
WS S P ( ZMV ) P
WP = with an internal standard
S s ( ZMV )s
90 Iron Ore

where WP is the relative mass fraction of phase p in a mixture of n phases, WP is the


absolute mass fraction of phase p in a mixture, WS is the mass fraction of standard s, SP
is the Rietveld scale factor for phase p, Ss is the Rietveld scale factor for the standard,
Z is the number of formula units per cell for phase i, M is the mass of the formula unit
for phase i, V is the cell volume for phase i.
If an amorphous phase is present and an internal standard was added, its mass
fraction is given as

Wis
1-
WRc
WAm =
100 - Wis

where WAm is the mass% amorphous phase, Wis is the weighed mass% standard, and
WRc is the Rietveld determined mass% of standard.
In some investigations, the full crystal structures of one or more phases are not
known, and Scarlett and Madsen (2006) had developed a method of quantification
(PONKCS), used in conjunction with the Rietveld method. In this case, the pure phase
or a known quantity of the phase is used to calibrate the method. The method shows
great promise even in cases where the crystal structures are well known, because effects
such as microabsorption that can seriously affect phase quantification are mitigated.
The Rietveld method matches the XRD pattern calculated from the crystal structure
data with the experimental pattern using a nonlinear least squares fitting technique.
The fit minimizes the following function:
n

( )
2
R = wi Yi obs - Yi calc
i =1

where Y obs and Ycalc are the observed and calculated intensities for every step of the
XRD pattern and w is a weighting factor.
The profile residual is given as
1/ 2
n
( )
2
wi Yi - Yi
obs calc

Rwp = i =1 n

( )
2
wi Yi
obs

i =1
A note of caution must be expressed here; often, samples containing transition ele-
ments exhibit high background counts, especially when modern strip detectors to-
gether with adjustable slits are used. This can give rise to low Rwp values because of
the predominant background (Y obs includes the background). This might therefore not
necessarily reflect a good fit.
Typical parameters that are obtained from a Rietveld refinement are the following:
Crystallographic information of the phases, that is, the crystal structures can be refined.
The lattice parameters of the phases. These can be used to calculate compositions in solid
solutions.
XRD analysis and evaluation ofiron ores and sinters91

Figure3.4 Rietveld analysis of an iron ore sinter showing the calculated pattern (red)
superimposed on the experimental pattern. The difference pattern is shown below the
calculated pattern of hematite. The peak positions of the different phases are also shown. The
quantities of the different phases are shown in the upper right-hand corner.

Relative quantities for the constituent phases.


Amorphous content if an internal standard is used.
Crystallite sizes of the phases.
Lattice strain of the phases. This is a rather broad term that includes stacking disorder, dis-
locations, and strain.

A typical Rietveld analysis of an iron ore sinter is shown in Figure 3.4. In this case,
the calculated pattern of the hematite phase is also shown together with the difference
curve (between experimental and calculated intensities). This should be horizontal and
featureless without any residual peaks.

3.4Sources of error in XRD analysis


Mineral quantification is the most widely used application of Rietveld analysis, and
it is applied to numerous cases, especially in geometallurgy and mineral processing.
However, despite its successes, the method still suffers from the following shortcom-
ings that can affect its accuracy and precision:
Preferred orientation of minerals with a pronounced habit such as micas and amphiboles.
The mineral grains orient themselves in the sample holder in response to their shape, and
therefore, the diffraction peaks from the preferentially oriented planes are artificially en-
hanced. Methods have been devised to correct for this effect but are still inadequate in some
cases.
92 Iron Ore

Microabsorption effects in samples containing phases with variable density. This effect is
due to the absorption of the diffracted X-rays by the minerals with bigger mass absorption
coefficients. This effect is especially prevalent if CuK radiation is used for the examination
of iron-containing minerals and phases.
XRD cannot easily detect or quantify minerals with less than ~1% abundance because of rel-
atively poor peak to background ratios. The method is therefore restricted to quantification
of ores containing major amounts of valuable minerals. Fortunately, iron ores and sinters fall
in this class, and therefore, XRD is particularly suited for their evaluation.
Quantitative phase analysis by XRD can also be seriously in error because of the presence
of undetected phases or minerals and the presence of amorphous material. These are by defi-
nition difficult to detect, and therefore, the results must always be validated by independent
means such as microanalysis of the phases and chemical analysis. This will be discussed later.

3.5Applications of cluster analysis


Cluster analysis is a statistical procedure that discriminates between sets of closely
related XRD scans from different samples. It then separates the similar samples into
different groups using statistical criteria. No information is given as to what the differ-
ences between groups entail: only that they are different. The procedure is especially
useful in ore body evaluation to discriminate between different ore types and in sinter
evaluation to examine differences between sinters prepared using different recipes or
operational variables such as suction and cooling rates. Once the different clusters
have been identified, they can then be examined in greater detail to establish what the
differences are and what they mean in terms of ore quality or sinter properties such as
tumble index or reduction disintegration indices. The principal components are com-
binations of features in the diffraction patterns, and no information is given regarding
what these components represent in terms of mineral identities or quantities. This
must be deduced by the investigator upon closer evaluation.
For an example of the application of cluster analysis to iron ore evaluation and
grade control, the reader is referred to the contribution by Paine etal. (2012), who
evaluated a large number of iron ore samples from an iron ore deposit.
In sinters, ongoing studies have identified differences according to different recipes
and also differences between size fractions, to be discussed later. Figure 3.5 shows the
differences between different clusters of sinter samples as small spheres of different
colors. For each of the clusters, the XRD patterns of the samples representing the clus-
ter are also given so that the characteristics of each cluster can be deduced more easily.

3.6Applicability of XRD analysis


The XRD method is not universally applicable to all minerals in ore deposits. This
is true for all ores where the economic minerals occur as minor or trace phases. Due
to their low signal to background ratio and the extensive overlap of the diffraction
peaks, they show relatively high detection limits as compared to elemental analysis.
However, for all ores where the economically important phases are the predominant
XRD analysis and evaluation ofiron ores and sinters93

8 +

13 ***

11 +
2.4
2.2 6 *** +
3 +
2
1.8 4 +
1.6
1.4
1.2 7 ***
5
Principal component 2

1
0.8
0.6
0.4
12 + 2
0.2
0 9 **
0.2 14
10 +
0.4
0.6
0.8 1
1 0.5
1.2 1
0
1.4
0.5 Principal component 3
1.6
1
1.8
1.5
3 2 1 0 1 2 3
Principal component 1
Figure3.5 Clustering of sinter samples using XRD patterns as input. The different clusters
are shown on a 3-dimensional plot. The biggest differences between samples are manifested
along principal component 1, which is a combination of a number of features in the diffraction
patterns.

ones, such as iron, aluminum, phosphate, and fluorite ores, XRD is widely used. Since
the method is one of the few that detect and quantify minerals, it can be used to mea-
sure the different minerals in iron ores. This can lead to better classification of ores
and eventually to causes of elemental variation in them.
In iron ore sinters, the presence of different phases reflects differences in sinter
constituent mixtures and different temperatures and reductant additions during sinter-
ing. This can be of use in the development of stronger, more reducible sinters of the
desired basicity.

3.7Use of mass balancing in iron ore analysis


Mass balancing is important in not only assessing the accuracy of XRD and mineral
analysis but also evaluating the suitability of ores for upgrading and beneficiation.
The bulk chemical composition is calculated from the XRD and mineral analyses
(by microprobe or SEM analysis) and is compared with the actual chemical analysis
94 Iron Ore

(by XRF or ICP analysis). This shows the contribution that each mineral makes to the
elemental abundances. A percentage deviation is calculated, which is an indication
of the accuracy of the analysis. Smaller deviations indicate a higher likelihood that
the XRD and phase compositions are correct. An example of such a calculation of a
high-alumina iron ore is given in Table 3.1.

3.8The principal minerals and phases


The minerals occurring in all major iron ores is given in Table 3.2, together with their
formulas and most important elemental impurities. These impurities, especially alu-
minum and magnesium in the main iron-bearing minerals, hematite, magnetite, and
goethite, are important factors to consider in the upgrading of low-grade iron ores.
In sinters, the main phases that form are listed in Table 3.3. Important are the main
impurities that affect the electron backscatter intensities of especially magnetite and
hematite, inhibiting their quantification using SEM-based methods.

3.9XRD for the characterization of Iron ores


3.9.1Iron ore characterization
Variations in iron ores can be traced and mapped using cluster analysis and XRD
quantification. Paine etal. (2012) evaluated a large number of iron ore samples from
an iron ore deposit. Using cluster analysis and mineral quantification, the ores could
be classified into defined theoretical grade blocks, which included high grade, high
grade with minor gibbsite, high-grade beneficiation, low-grade beneficiation, low-
grade other, and waste. As a result, material with a propensity for higher degrees of
beneficiation was identified and delimited.
For iron ore beneficiation, the mineral quantities in the ores is essential to establish
the degree of upgrading that can be achieved. In a study of the removal of aluminum in
goethitic iron ores, mass balance calculations assisted greatly to assess the maximum
amount of Al that can be removed without appreciable iron loss, mainly from the
goethite. This is shown graphically in Figure 3.6, which shows that 68% of the Al in
the sample is distributed in goethite. The goethite also contains 60% of the iron in the
sample and cannot be removed. Therefore, if Al is to be removed, only kaolinite and
gibbsite can be eliminated without major iron loss, and only as little as 22% of the Al
can be removed by flotation or other methods.

3.9.2Lattice constant refinement for the evaluation


of hematite and goethite
Lattice constant refinement can be used to assess the substitution of impurity ele-
ments, especially in fine-grained goethite and hematite, as determined by Schulze
(1984) and Stanjek and Schwertmann (1992), respectively.
XRD analysis and evaluation ofiron ores and sinters95
Table3.1 Mass balance comparing XRD and mineral analysis
Mass balance calculationFe ore FCF11-A

Mass balance Fraction(XRD) SiO2 Al2O3 TiO2 Fe2O3 LOI P2O5 K2O
Annite 0.21 0.06 0.02 0.00 0.09 0.00 0.00 0.03
Gibbsite 0.17 0.00 0.11 0.00 0.00 0.06 0.00 0.00
Goethite 60.48 2.49 3.09 0.81 47.80 6.29 0.00 0.00
Hematite 32.78 0.33 0.46 0.00 31.99 0.00 0.00 0.00
Kaolinite 2.17 1.00 0.85 0.00 0.00 0.30 0.00 0.00
Magnetite 3.82 0.00 0.01 0.00 3.94 0.00 0.00 0.00
Quartz 0.38 0.38 0.00 0.00 0.00 0.00 0.00 0.00
Total 100.01 4.27 4.54 0.81 83.83 6.64 0.00 0.04 100.12 R(%)
XRF Analysis 3.59 4.54 0.69 82.58 7.44 0.11 0.00 98.95 3.03
96 Iron Ore

Table3.2 Principal minerals occurring in sedimentary iron ores


Mineral name Formula Impurities
Hematite Fe2O3 Al2O3 and OH
Magnetite Fe3O4 MgO and Al2O3
Goethite FeOOH Al2O3 and adsorbed P2O5
Greenalite (Fe2+,Fe3+)23Si2O5OH4 MgO
Stilpnomelane K(Fe2+,Mg,Fe3+)8(Si,Al)12
(O,OH)27n(H2O)
Kaolinite Al2Si2O5(OH)4 Almost pure
Annite KFe32+AlSi3O10(OH,F)2
Gibbsite Al(OH)3 Almost pure
Quartz SiO2 Pure

Table3.3 Predominant phases in iron ore sinters


Phase Formula Impurities
Hematite Fe2O3 Al2O3
Magnetite Fe3O4 CaO, MgO, and Al2O3
SFCA (silico-ferrite of Ca (Ca,Mg)3(Fe,Al)10SiO20 MnO
and Al)
SFCA-I Ca3(Fe,Al)17O28 MnO
SFCA-II Ca5(Fe,Al)29O48 MnO
Larnite (C2S) Ca2SiO4 FeO

FCF11-A Al distribution FCF11-A Fe distribution


Annite Gibbsite
Kaolinite Gibbsite 0% 0%
19% 3%
Kaolinite
0%
Hematite Hematite
10% 40%
Goethite Goethite
68% 60%

Figure3.6 Distribution of Al and Fe in the different phases in a goethite-rich iron ore.

The hematite unit cell parameters are affected by Al and OH substitution:

a = 5.0359 - 0.00183Al + 0.00175LOI

c = 13.740 - 0.00512 Al + 0.0130 LOI


XRD analysis and evaluation ofiron ores and sinters97

Table3.4 Calculation of Al and OH contents in hematite and Al


contents in goethite using the unit cell parameters of the two
phases
Hematite Goethite

Sample X(Al)-h X (OH) LOI(%) c Mol % Al X(Al)-g


FCF 11-A 0.039 0.023 0.42 3.0154 5.22 0.052
FCF 11-B 0.022 0.042 0.78 3.0187 3.28 0.033
FCF 11-C4 0.032 0.054 1.01 3.0201 2.50 0.025
FCF 11-D4 0.027 0.034 0.62 3.0180 3.73 0.037
FCF 11-D4 0.058 0.042 0.78 3.0178 3.80 0.038

OH substitution affects mainly the c parameter:

Dc = 0.00841 + 0.0130 LOI

The goethite cell parameters are affected mainly by Al substitution:

%Al = 1730 - 572 * c

The procedure was elegantly included in the Rietveld refinement procedure by


Knorr and Neumann (2011) and Neumann etal. (2014).
The use of XRD can therefore give a quick assessment of the extent of Al and OH
substitution in hematite and the amount of Al substitution in goethite. This was done
for five goethite-rich iron ores and is shown in Table 3.4.

3.10XRD in sintering and pelletizing


3.10.1Fundamental studies of sinter phases and reactions
Reliable quantification using Rietveld analysis rely critically on accurate crystal struc-
tures of minerals and sinter phases. Fortunately, the structures of the important binding
phases SFCA, SFCA-I, and SFCA-II have been determined by Mumme and coworkers
(Hamilton etal., 1989; Mumme etal., 1998; Mumme, 2003). Sugiyama etal. (2005)
determined the structures of Mg-rich SFCA. Recently, the crystal structures of SFCA
and aluminium free SFC have been redetermined by Liles and de Villiers (2015).

3.10.2Sinter reactions
Important laboratory studies of reactions in sinter mixtures have been conducted at
high temperature at controlled oxygen partial pressures using both laboratory and syn-
chrotron X-ray sources by Scarlett etal. (2004) and Webster etal (2012). Critical
reactions and formation temperatures of the various phases have been established.
98 Iron Ore

These serve as important benchmarks for the interpretation of the reactions taking place
in course-grained mixtures in experimental sinter pots and commercial sinter plants.
Reactions that take place during sintering and pelletizing can be easily followed by
XRD. The extent of sintering can be assessed by the SFCA/hematite ratio; this is not
possible by using chemical analysis. The redox conditions prevailing during sintering
(reductant addition) can be assessed by the magnetite/hematite ratio. Presently, this
is determined chemically using the FeO content. The relation between the XRD peak
intensities and the FeO content was established by Bonvin etal. (2000) and by Konig
etal. (2012).
The reactions that occur during initial combustion, at maximum temperature, and
during cooling can be established by interrupting the sinter reactions in a sinter box
using water quenching. Figure 3.7 shows the diffraction patterns of samples collected
from top to bottom in a quenched sinter pot. This shows that hematite at the bottom of
the sinter pot essentially disappears at the flame front (F4 and F5) with the formation
of magnetite and that secondary hematite (and SFCA) again forms during cooling and
oxidation by air at the top of the sinter bed.

3.10.3Phase chemistry of sinter screen fractions


Figure 3.5 shows an essential difference between clusters along principal component 1.
This difference is between size fractions of sinters. Hematite is concentrated in the sam-
ples of 5mm fractions that cluster on the left side of the diagram. This is also shown in
Figure 3.8. The production of sinter fines is probably due to the formation of secondary
or granular hematite, a product that contributes to the degradability of sinters.

Counts
4000 2WNRF1
2000
0
2WNRF2
4000
0
4000 2WNRF3
2000
0
2WNRF4
4000
0
2WNRF5
4000
0
2WNRF6
4000

0
10 20 30 40 50 60 70 80
Position (2q ) (Cobalt (Co))

Peak List
Fe3O4
Fe2 O3
Ca2.3 Mg0.8 Al1.5 Fe8.3 Si1.1 O20
Ca3.18 Fe15.48 Al1.34 O28

Figure3.7 Changes in the phase composition from top to bottom in a water-quenched


sinter pot.
XRD analysis and evaluation ofiron ores and sinters99

HEM

5 mm fraction

12.5 + 10 mm
fraction

SFCA MAG
Figure3.8 Triangular plot showing that the 5mm fraction is enriched in hematite.

3.11Summary and conclusions


The use of XRD and the Rietveld method have resulted in the development of fast
and reliable methods of phase quantification in the iron ore industry. Compared with
other methods, it is still the most reliable method for phase quantification, and it has
the potential to make important contributions to our knowledge regarding the effect of
phase relations on the properties of sinters. Several applications of the method have
been discussed in some detail.
New developments will necessarily attempt to relate phase composition to sinter
strength and reducibility, and increased use of cluster analysis can identify important
relations between these parameters.

References
Bonvin, D., Yellepeddi, R., Buman, A., 2000. Applications and perspectives of a new innovative
XRF-XRD spectrometer in industrial process control. Adv. X-Ray Anal. 42, 126136.
Hamilton, J.D.G., Hoskins, B.F., Mumme, W.G., Borbridge, W.E., Montague, M.A., 1989. The
crystal structure and crystal chemistry of Ca2.3Mg0.8Al1.5Fe9.3O20 (SFCA): solid solution
limits and selected phase relationships of SFCA in the SiO2-Fe2O3-CaO(-Al2O3) system.
N. Jb. Mineral. (Abh.) 161, 126.
100 Iron Ore

Hill, R.J., Howard, C.J., 1987. Quantitative phase analysis from neutron powder diffraction data
using the Rietveld method. J. Appl. Crystallogr. 20, 467474.
Jenkins, R., Snyder, R.L., 1996. Introduction to X-ray Powder Diffractometry. John Wiley,
New York.
Klug, H.P., Alexander, L.E., 1974. X-ray Diffraction Procedures for Polycrystalline and
Amorphous Materials, second ed. John Wiley, New York.
Knorr, K., Neumann, R., 2011. Advances in quantitative X-ray mineralogymixed crystals
in bauxite. In: Proceedings of the 10th International Congress for Applied Mineralogy
(ICAM), TRONDHEIM, Norway, pp. 377384.
Konig, U., Gobbo, L., Reiss, C., 2012. Quantitative XRD for ore, sinter, and slag characteriza-
tion in the steel industry. In: Proceedings of the 10th International Congress for Applied
Mineralogy (ICAM), pp. 385393.
Liles, D.C., de Villiers, J.P.R., 2015. Refinement of iron ore sinter phases: a silico-ferrite of
calcium and aluminium (SFCA) and an Al-free SFC, and the effect on phase quantification
by X-ray diffraction. Acta Crystallogr. accepted for publication.
Madsen, I.C., Scarlett, N.V.Y., Cranswick, L.M.D., Lwin, T., 2001. Outcomes of the International
Union of Crystallography Commission on powder diffraction Round Robin on quantitative
phase analysis: samples 1a to 1h. J. Appl. Crystallogr. 34 (4), 409426.
Mumme, W.G., 2003. The crystal structure of SFCA-II, Ca5.1Al9.3Fe3+18.7Fe2+0.9O48, a new
homologue of the aenigmatite structure type, and structure refinement of SFCA type,
Ca2Al5Fe7O20. Implications for the nature of the ternary phase solid solution previously
reported in the CaO-Al2O3-iron oxide system. N. Jb. Mineral. (Abh.) 178, 307335.
Mumme, W.G., Clout, J.M.F., Gable, R.W., 1998. The crystal structure of SFCA-I, Ca3.18
Fe3+14.66Al1.34Fe2+0.82O28, a homologue of the aenigmatite structure type, and new crystal
structure refinements of -CFF, Ca2.99Fe3+14.30Fe2+0.55 O25 and Mg-free SFCA, Ca2.45Fe3+9.04
Al1.74Fe2+0.16O20. N. Jb. Mineral. (Abh.) 173, 93117.
Neumann, R., Avelar, A.N., da Costa, G.M., 2014. Refinement of the isomorphic substitutions
in goethite and hematite by the Rietveld method, and relevance to bauxite characterisation
and processing. Miner. Eng. 55, 8086.
Paine, M., Knig, U., Staples, E., 2012. Application of rapid X-ray diffraction (XRD) and
cluster analysis to grade control of iron ores. In: Proceedings of the 10th International
Congress for Applied Mineralogy (ICAM), pp. 495501.
Pecharsky, V.K., Zavalij, P.Y., 2005. Fundamentals of Powder Diffraction and Structural
Characterization of Materials. Springer, New York.
Scarlett, N.V.Y., Madsen, I.C., 2006. Quantification of phases with partial or no known crystal
structures. Powder Diffract. 21 (4), 278284.
Scarlett, N.V.Y., Madsen, I.C., Pownceby, M.I., Christensen, A.N., 2004. In situ X-ray diffrac-
tion analysis of iron ore sinter phases. J. Appl. Crystallogr. 37, 362368.
Schulze, D.G., 1984. The influence of aluminium on iron oxides. VIII. Unit cell dimensions of
Al-substituted goethites and estimation of Al from them. Clays Clay Miner. 32 (1), 3644.
Stanjek, H., Schwertmann, U., 1992. The influence of aluminium on iron oxides. Part XVI: hy-
droxyl and aluminium substitution in synthetic hematites. Clays Clay Miner. 40 (3), 347354.
Sugiyama, K., Monkawa, A., Sugiyama, T., 2005. Crystal structure of the SFCAM phase.
Ca2(Ca,Fe,Mg,Al)6(Fe,Al,Si)6O20. ISIJ Int. 45 (4), 560568.
Webster, N.A.S., Pownceby, M.I., Madsen, I.C., Kimpton, J.A., 2012. Silico-ferrite of calcium
and aluminum (SFCA) iron ore sinter bonding phases: new insights into their formation
during heating and cooling. Metall. Mater. Trans. B 43B, 13441357.
Young, R.A. (Ed.), 1993. The Rietveld method. In: IUCr Monographs on Crystallography.
Oxford University Press, Oxford, New York.

S-ar putea să vă placă și