You are on page 1of 12

Chemical Geology 231 (2006) 288 – 299

www.elsevier.com/locate/chemgeo

Structural characterization and chemical reactivity of synthetic


Mn-goethites and hematites
Mariana Alvarez a,1 , Elsa H. Rueda a,⁎, Elsa E. Sileo b
a
Departamento de Química, Universidad Nacional del Sur, Av. Alem 1253, B8000CPB, Bahía Blanca, Argentina
b
INQUIMAE, Dpto. de Química Inorgánica, Analítica y Química Física, Facultad de Ciencias Exactas y Naturales, Universidad de Buenos Aires,
Pabellón II, Ciudad Universitaria, C1428EHA, Buenos Aires, Argentina
Received 30 November 2005; accepted 1 February 2006

Abstract

Several goethites were obtained through the hydrolysis at 60 °C of Fe(III) solutions containing variable amounts of Mn(II) ions.
The obtained samples were thermally treated at temperatures ranging from 180 to 310 °C until the complete phase transformation
to hematite was achieved. The effect of Mn in the dehydroxylation process was investigated using X-ray diffraction (XRD) and the
Rietveld refinement of XRD data together with scanning electron microscopy (SEM), differential thermogravimetric analysis
(DTA) and Fourier transform infrared spectroscopy (FTIR). In all cases, the formed hematites retained the acicular shape of the
precursor goethite. The dehydroxylation temperature increased with the increase of the Mn content in the parent goethite. The cell
parameters of both phases decreased with the thermal treatment, however the decrease in the goethite b-parameter was more
pronounced. This fact could be attributed to the distortion in the goethite structure by the presence of manganese. The band shifts in
the FT-IR spectra of the goethites with different Mn substitution were analysed. The intensities of the hydroxyl vibrations were
indicative of the degree of dehydroxylation.
The chemical reactivity of all the samples, before and after the thermal treatment, was also studied. The kinetic experiments
were carried out at 40 °C in 4 mol dm− 3 HCl. The acid dissolution of all Mn-goethites showed a congruent behavior indicative of a
homogeneous distribution of Mn in the goethite crystals, this trend was not observed in the formed hematites presenting a high Mn
content. The dissolution rate in goethites increased with the increase of Mn content, the opposite effect was observed in the
corresponding hematites. The activation energy in both phases was also obtained and indicated that the Mn substitution produces
an opposite effect on goethite- and hematite-phases. Different kinetic laws were applied in order to explain the dissolution behavior,
but the modified first-order Kabai equation described the dissolution data best.
© 2006 Elsevier B.V. All rights reserved.

Keywords: Mn-goethite; Thermal transformation; Mn-hematite; Rietveld refinement; Acid dissolution

1. Introduction due to the wide interest in the technological applications


of iron oxides. A common method of hematite formation
Several studies involving different aspects of the is the thermal dehydration of goethite; this has been
dehydration of goethite to hematite have been performed the focus of a relative large number of studies (Derie
et al., 1976; Klissurski and Bluskov, 1980; Mendelovici
⁎ Corresponding author. Tel.: +54 291 4595159. et al., 1982; Goñi-Elizalde and García-Clavel, 1988;
E-mail address: ehrueda@criba.edu.ar (E.H. Rueda). Schwertmann and Cornell, 1991; Pomiès et al., 1999).
1
Tel.: +54 291 4595159. The final product of the thermal dehydration is α-Fe2O3,
0009-2541/$ - see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.chemgeo.2006.02.003
M. Alvarez et al. / Chemical Geology 231 (2006) 288–299 289

whose properties depend on the temperature, treatment information regarding the distribution of the foreign
conditions, and the crystallinity of the goethite (Derie et metal within the structure of the iron oxide. If Fe and
al., 1976; Klissurski and Bluskov, 1980; Goñi-Elizalde associated metals dissolve at identical rates, then it is
and García-Clavel, 1988). Dehydration of goethite in- reasonable to assume that the metal is uniformly dis-
volves removal of hydrogen and one quarter of the tributed throughout the iron oxide crystals. Conversely,
oxygen, without disturbing the network of remaining if Fe and the metal show quite different patterns of
oxygen, and atomic rearrangement of Fe3+ to form dissolution, then it is unlikely that the two elements are
hematite: 2 α-FeOOH → α-Fe2O3 + H2O. Two different present in a single mineral of uniform composition
mechanisms have been suggested: direct dehydration of (Sidhu et al., 1981; Singh and Gilkes, 1992).
goethite to hematite without an intermediate phase The aim of this work is to provide information about
(Watari et al., 1983; Goss, 1987), and the formation of the effects of Mn substitution on the thermal transfor-
an intermediate product during dehydration where in the mation of goethite to hematite achieved at different
inhomogeneous dehydration process, the crystal devel- temperatures. For that, the structural changes of the two
ops separate ‘donor’ and ‘acceptor’ regions (Ball and phases are analysed using the Rietveld refinement of
Taylor, 1961; Lima-de-Faria, 1963). XRD powder data. The effect of the Mn incorporation
Although most work on the goethite dehydroxylation on the dissolution behavior in acidic media of all the
has focused on non-substituted goethite (Ruan et al., formed solids is also investigated.
2001), natural goethite often contains appreciable quan-
tities of aluminum ion substituting Fe. This substitution 2. Materials and methods
may affect the dehydroxylation temperature (Jonas and
Solymar, 1970; Fey and Dixon, 1981; Ruan and Gilkes, Manganese goethites were synthesized from the
1995a). Others cations such as Cr, Ge, Ni, Co and Mn Fe(III)-system using a method described elsewhere
may also substitute Fe in natural goethite (Schwertmann (Sileo et al., 2001). Ferrihydrites were precipitated by
and Lathman, 1986; Bernstein and Waychunas, 1987; adding a 2 mol dm− 3 NaOH to different solutions
Trolard et al., 1995). containing variable quantities of Fe(III) and Mn(II)
Chemical dissolution of iron oxides has been studied nitrates. The final ratio OH− / (Fe + Mn) and the total
for a long time in analysis of natural and synthetic (Fe + Mn) concentration in all samples were 13.16 and
mixtures, material chemistry, and technology (Blesa et 0.53 mol dm− 3, respectively. The precipitates obtained,
al., 1994; Cornell and Schwertmann, 1996). In general, were washed twice with double-distilled water and
the solubility of iron oxides in pure water is extremely centrifuged. After this treatment, the samples were aged
low (Schwertmann, 1991); however, dissolution of Fe for 21 d at 60 °C in closed polyethylene flasks con-
oxides is promoted by strong acids and by reducing and taining 0.3 mol dm− 3 NaOH. Initial mole ratios
complexing agents. χMn = Mn / (Mn + Fe) were 0.00, 0.03, 0.07 and 0.10
Dissolution of iron oxides containing foreign ions (samples named S0_60 °C to S3_60 °C, respectively). After
that replace Fe has also received attention. The mech- aging, the materials were washed with double-distilled
anisms of dissolution of pure and Al-substituted, syn- water until the conductivity of the filtrated solution was
thetic and natural goethite in HCl solutions are well b 5 μS. The remaining solids were dried at 40 °C and
understood (Surana and Warren, 1969; Cornell et al., gently crushed. Then, the samples were oxalate-ex-
1974, 1975, 1976; Schwertmann, 1984a; Cornell and tracted in order to remove remaining ferrihydrite from
Schindler, 1987). Lim-Nuñez and Gilkes (1987) and their surfaces.
Alvarez et al. (2005) showed that transition metals- Sub-samples of Mn-substituted goethites were heated
substituted goethites dissolve in acidic media at rates in a glass furnace with temperature control within ± 2 °C
different from that of pure goethite. Wells et al. (2001) for 2 h at temperatures ranging from 180 to 310 °C
investigated the influence of metal substitution on the to evaluate the degree of transformation goethite →
dissolution rate and kinetics of dissolution in 1 mol dm− 3 hematite.
HCl of Al-, Mn- and Ni-susbtituted hematites. In all The level of Mn incorporation was calculated from
cases, dissolution kinetics have been used to interpret the composition of the oxalate-extracts. Metal contents
the release mechanism of cations included in goethite in the goethites were determined on 20 mg of each
and hematite; so, the acid dissolution provides evidence sample dissolved in 6 mol dm− 3 HCl at 60–80 °C, using
of the uniformity of incorporation of Fe and metals in a GBC Model B-932 flame atomic absorption spectrom-
these minerals. Iron and metal analysis of sequential eter. Differential thermal analyses (DTA) were carried
extracts during the dissolution of samples can provide out in a DTA-50 Shimadzu analyzer. Approximately
290 M. Alvarez et al. / Chemical Geology 231 (2006) 288–299

Table 1 ultrasonic treatment and a drop of suspension was placed


Chemical composition for the samples synthesized at 60 °C onto a metallic support. SEM images were obtained from
Sample(1) χMn (preparative) χMn (experimental) a Phillips SEM 515 equipment operated at 30 keV.
S0_60 °C 0.00 0.0 Fourier Transform infrared absorption (FTIR) spec-
S1_60 °C 0.03 0.025 ± 0.005 tra, between 400 and 4000 cm− 1 with 2 cm− 1 resolution,
S2_60 °C 0.07 0.067 ± 0.001 were obtained using a Nicolet 510P FT-IR spectrometer.
S3_60 °C 0.10 0.083 ± 0.002
The samples were dispersed in KBr disks (1%).
(1)
Subscript indicates different Mn level and synthesis temperature. Dissolution kinetics were measured at 40 °C. 50 mg
of each sample was suspended in 50 cm3 of 4 mol dm− 3
5 mg of sample was heated in flowing air to 500 °C at HCl and sealed in a cylindrical beaker provided with a
10 °C min− 1. thermostat water jacket. The suspension was magneti-
X-ray diffraction patterns were recorded in a Siemens cally stirred throughout the experiment. Sample aliquots
D5000 diffractometer using the Cu Kα radiation at (1 cm3) were withdrawn at predetermined intervals and
40 kV, and 30 mA. Divergence, scattered, and receiving filtered immediately through 0.22 μm membrane filters.
radiation slits were 1°, 1°, and 0.2 mm, respectively. A The solution to solid weight ratio (1000 : 1) used in this
curved graphite-monochromator was used. Data were work is in line with previous dissolution studies that
collected in the 19.000° ≤ 2θ ≤ 140.000° range, with used ratios in the range 250–2000 : 1 to ensure that the
variable scanning steps in the range 0.030° to 0.020° dissolution was independent of the solution to sample
and different counting times. The step width of each run ratio (Cornell et al., 1974; Sidhu et al., 1981; Lim-Nuñez
assured a minimum of about 12 intensity points for the and Gilkes, 1987). Determinations of dissolved Fe and
narrower peaks. The data were analysed by the Rietveld
refinement method (Rietveld, 1969) using the General
Structure Analysis System (Larson and Von Dreele,
1994). The Thompson–Cox–Hasting pseudo-Voigt
function (Thompson et al., 1987) with the microstrain
broadening description of Stephens (1999) was used for
fitting the peak profiles. A linear interpolation function
was employed for fitting the background signal. Cell
parameters, peak displacement, full width at half
maximum, microstrain broadening, scale factor, and
positional parameters for all atoms were refined. The
isotropic thermal parameters were fixed at the values
measured for goethite (Szytula et al., 1968) and hematite
(Maslen et al., 1994) single-crystals.
Specimens for scanning electronic microscopy
(SEM) were dispersed in double-distilled water with

Fig. 1. DTA profiles for the samples Si_60 °C. Mn contents are indicated Fig. 2. XRD of samples (a) S0 and (b) S3 heated at different
in Table 1. temperatures.
M. Alvarez et al. / Chemical Geology 231 (2006) 288–299 291

In addition, a splitting of the dehydroxylation endo-


thermic peak is observed. A double DTA peak due to
dehydroxylation of pure goethite to hematite has been
found by Derie et al. (1976) and Schwertmann (1984b).
Derie et al. (1976) suggested that the hematite shell
formed around goethite crystals retards the dehydration
of the core, while Murad reported that ground natural
highly crystalline goethite shows a double peak.
Schwertmann (1984b) also reported a double peak in
the dehydroxylation of well-crystallized α-FeOOH that
was attributed to an intermediate goethite with a slightly
higher dehydroxylation temperature generated during
heating. In Al-substituted goethite the double peak was
interpreted by the difference in the dehydroxylation
temperatures of Al–OH and Fe–OH groups (Fey and
Dixon, 1981). However, these explanations for a double
peak cannot be used to explain the multiple peaks
observed here. Instead polycrystallinity and lattice dis-
tortions and defects may be responsible for the peak
broadening. The presence of Mn in goethite weakens
and broadens this endothermic feature.
The XRD patterns (Fig. 2) obtained after heating the
samples at different temperatures show that goethite is
completely transformed to hematite from 260 to 310 °C
as the χMn value increases from 0 to 0.083 mol/mol.

3.2. Crystal morphology

Fig. 3. Scanning electron micrographs (×20,000) of samples


Fig. 3 shows SEM micrographs for sample S2 at three
(a) S2_60 °C and (b) S2_310 °C. different heating temperatures. The starting goethite
crystals prepared at 60 °C are acicular and this
Mn amounts were determined by atomic absorption morphology is retained on thermal treatment at 220
spectroscopy.
To determine activation energy (Ea), dissolution Table 2
kinetics were measured at three temperatures (40, 50 Agreement factors for the refinements
and 60 °C). Sample(1) R(2)
wp R(3)
exp GofF(4) RBragg Composition (%)
All kinetic experiments were carried out in duplicate Goethite Hematite
or triplicate and the coefficient of variation was b 0.08. S0_60 °C 10.50 7.24 1.4 4.21 100.0 0.0
S1_60 °C 10.50 7.27 1.4 3.78 100.0 0.0
3. Results and discussion S2_60 °C 10.60 7.00 1.5 3.67 100.0 0.0
S3_60 °C 8.98 6.95 1.3 2.91 100.0 0.0
S0_220 °C 11.18 8.88 1.50 7.55 58.6(8) 41.4(8)
3.1. Mn content, DTA and XRD
S1_220 °C 9.15 7.18 1.38 3.91 61.4(4) 38.6(4)
S2_220 °C 8.83 6.67 1.49 3.51 68.1(4) 31.9(4)
Table 1 shows that the Mn content in the series S3_220 °C 11.08 8.60 1.18 5.15 71.4(3) 28.6(3)
increases with the increase of Mn (II) in the initial S0_260 °C 9.07 7.13 1.42 4.96 0.0 100.0
solution. S1_260 °C 9.35 7.27 1.51 4.01 0.0 100.0
S2_260 °C 9.92 7.50 1.57 3.83 30.4(8) 69.6(8)
The main endothermic DTA peak (Fig. 1) for the
S3_260 °C 11.63 9.19 1.55 9.46 35.9(7) 64.1(7)
dehydroxylation is displaced from 267 to 294 °C as the S2_310 °C 13.81 10.42 1.40 8.88 0.0 100.0
Mn content increases from χMn = 0 to 0.083. The same S3_310 °C 12.47 9.93 1.36 6.04 0.0 100.0
trend was observed for Al-(Fey and Dixon, 1981; Lim- (1)
Subscript indicates dehydroxylation temperature; (2)Rwp: weighed
Nuñez, 1985; Ruan and Gilkes, 1995a) and Cr- sum of residuals of the least square fit of the Rietveld refinement;
(3)
substituted goethites (Sileo et al., 2004). Rexp: value statistically expected; (4)GofF: Rwp / Rexp.
292 M. Alvarez et al. / Chemical Geology 231 (2006) 288–299

transformation of goethite to hematite was observed in


S0_260 °C and in S1_260 °C. At 310 °C only hematite was
present in all compositions (Fig. 4). The data confirm
that Mn-substitution in goethite raises the dehydroxyla-
tion temperature.
The cell parameters obtained by the Rietveld refine-
ment are shown in Table 3.
Although Fe(III) and Mn(III) have similar ionic
radii, the electronic configuration of Mn(III), a d4 ion,
presents a tetragonally distorted rather than a symmet-
rical octahedral coordination sphere that does not
arrange as easily into the structure of goethite as does
Fig. 4. Weight (%) of hematite from S0 to S3 heated at different Fe(III). This is reflected in the increase in the b
temperatures between 60 and 310 °C.
parameter as the level of Mn substitution for samples
synthesized at 60 °C increases and a decrease in the a
and 310 °C. The same feature was also observed by and c parameters (Sileo et al., 2001). These changes in
Cornell and Schwertmann (1996), Pomiès et al. (1999) cell parameters confirm the oxidation of Mn(II) to Mn
and González et al. (2000). (III) and the isostructural Mn-for-Fe substitution.
Fig. 5 shows the evolution of the cell parameters of
3.3. Hematite / goethite ratio and unit cell dimensions goethite and hematite vs. temperature for S2 and S3
during dehydroxylation samples.
A systematic decrease in the a and c parameters of
Table 2 collects the parameters that describe the Mn-goethite on thermal treatment takes place in both
quality of the Rietveld simulation and the proportion of samples; b parameter shows the same decreasing trend,
goethite and hematite in the samples. Reliability factors but in the temperature range 60–220 °C this change is
are in the range Rwp ≡ 8.83–13.81; RB ≡ 2.91–9.46; more pronounced in sample S3. At 260 °C both samples
GofF values are reasonably adequate, between 1.30 and reach practically the same value for this parameter. This
1.57. behavior of the b dimension may be related to the
Since the thermal treatment of samples S0 to S3 at 60 migration of cations and rearrangement of oxygen
and 180 °C renders only goethite, only refinement data packing over the temperature range of dehydroxylation
for 60 °C is presented. Partial dehydroxylation at 220 °C used in this study. Similar trend is observed in Al-
is occurred in all samples, and in samples S2 and S3 at goethites (Ruan and Gilkes, 1995a). It is important to
260 °C. For each temperature the higher goethite content emphasize that the Mn distribution within the goethite
is found in samples with higher χMn values. Complete and hematite crystals could not be determined. Gasser et

Table 3
Unit cell parameters for the samples
Sample Goethite Hematite
3
a (nm) b (nm) c (nm) Vol (nm ) a = b (nm) c (nm) Vol (nm3)
S0_60 °C 0.46172(5) 0.99548(10) 0.30242(3) 0.13901(2) – – –
S1_60 °C 0.46113(2) 0.99573(5) 0.30217(1) 0.13875(1) – – –
S2_60 °C 0.46051(1) 0.99698(2) 0.30197(1) 0.13864(1) – – –
S3_60 °C 0.45990(1) 0.99837(1) 0.30165(1) 0.13851(1) – – –
S0_220 °C 0.46114(9) 0.99774(7) 0.30051(4) 0.13827(4) 0.50519(63) 1.37576(262) 0.30408(99)
S1_220 °C 0.46060(10) 0.99791(12) 0.29902(5) 0.13744(3) 0.50553(20) 1.37905(70) 0.30521(16)
S2_220 °C 0.46066(9) 0.99686(12) 0.29819(5) 0.13693(3) 0.50632(11) 1.37893(179) 0.30615(39)
S3_220 °C 0.45945(5) 0.99540(8) 0.29762(2) 0.13611(2) 0.50481(25) 1.38673(72) 0.30604(24)
S0_260 °C – – – – 0.50390(2) 1.37624(13) 0.30263(3)
S1_260 °C – – – – 0.50373(5) 1.37583(20) 0.30233(7)
S2_260 °C 0.45588(18) 0.99546(38) 0.29418(13) 0.13350(5) 0.50467(5) 1.37375(37) 0.30300(8)
S3_260 °C 0.45832(10) 0.99559(18) 0.29465(4) 0.13444(4) 0.50312(4) 1.37499(21) 0.30142(5)
S2_310 °C – – – – 0.50333(2) 1.37537(18) 0.30176(4)
S3_310 °C – – – – 0.50308(4) 1.37535(16) 0.30145(6)
M. Alvarez et al. / Chemical Geology 231 (2006) 288–299 293

Fig. 5. Cell parameters vs. temperature for S2 (○) and S3 (■): (a), (b), (c) goethite phase, and (d), (e) hematite phase.

al. (1999) found a Mn zoning in goethite, which may 3.4. FT-IR spectra
partly explain inhomogeneous transformation of ferri-
hydrite to goethite. This behavior could also occur in the In the FT-IR spectra (Fig. 6) structural OH of
topotatic goethite–hematite phase transformation in the goethite causes the absorption at 3140 cm− 1, the
presence of Mn. hydroxyl stretching region, and the two bands at
The a and c parameters of hematite decrease as the ∼890 and 795 cm− 1 correspond to OH deformation
temperature of the treatment is raised from 220 to and water bending modes (Cambier, 1986). The
intensity of the band at 3140 cm− decreased as the
1
260 °C. Similar trends were reported for pure hematites
and Al-hematites obtained from thermal transformation Mn content increased shifting to lower energy values and
of the corresponding goethites (Ruan and Gilkes, 1995a; disappeared in all the samples after heating at 260 °C.
Ruan et al., 2001). The same trend in intensity is found in the bands
294 M. Alvarez et al. / Chemical Geology 231 (2006) 288–299

Fig. 6. Infrared spectra of samples (a) S0, (b) S1, (c) S2 and (d) S3, and its dehydroxylated products obtained from heating goethite between 180 and
310 °C for 2 h.

corresponding to the OH deformation and water bending Dissolution curves for goethite show nearly linear
modes. shapes in the range of time studied in this work
After heating at 220 °C all samples show a band at (Fig. 7a), whereas the curves for the samples S0_260 °C,
∼3400 cm− 1, together with the weak absorption at S1_260 °C, S2_310 °C and S3_310 °C, where the unique
∼1620 cm− 1, attributed to ‘loosely bound water’, which phase is hematite, show deceleratory shapes (Fig. 7b).
are indicative of hydroxyl in hematite. The spectra show As shown in the Fig. 7, dissolution rates are similar for
that the hematites, formed by thermal dehydroxylation S0_60 °C and S1_60 °C and then increase with Mn content
of goethite, inherited some structural OH expressed by (S0_60 °C ≈ S1_60 °C b S2_60 °C b S3_60 °C), whereas the
Fe2−x / 3(OH)xO3−x. Wolska and Szajda (1985) reported opposite effect was observed with the increase of Mn
the bands near 3400, 950 and 630 cm− 1 as OH unit substitution in the hematite samples. These latter re-
vibrations in “hydrohematite”. sults agreed with those presented by Wells et al. (2001).
The presence of goethite detected by XRD in The dissolution rate per gram of oxide was calculated
samples S2 and S3 only could be supported by FT-IR from the initial nearly linear region of the dissolution
spectra of S2. In samples S2_310 °C and S3_310 °C, the curves (Table 4).
spectra show characteristic bands of “hydrohematite”, The initial dissolution rates in Table 4 indicate that
and no bands of goethite are detected. These results non-substituted hematite (S0_260 °C) dissolved about 90
agree with XRD patterns. times faster than the corresponding goethite (S0_60 °C).
In the case of substituted samples, hematite with higher
3.5. Dissolution curves Mn content (S3_310 °C) dissolved only about 5 times
faster than the corresponding Mn-goethite (S3_60 °C).
The representative plots of Fe released at 40 °C in Then, Mn substitution reduced the dissolution rate of
4 mol dm− 3 HCl vs. time for the prepared Mn-goethites hematite, but increased the dissolution rate of goethite.
and the totally thermal transformed Mn-hematites are The increase in k0 values detected in the mixed
shown in Fig. 7. goethite–hematite samples compared to that of pure
M. Alvarez et al. / Chemical Geology 231 (2006) 288–299 295

goethite may be explained by the development of Table 4


micropores and structural disorder formed during the Initial dissolution rate (k0) for goethites as affected by Mn substitution
and heating
phase transformation (Naono and Fujiwara, 1980; Goss,
Sample Mol% Mn Hm (1) k0 (mmol Fe g R2
1987; Ruan and Gilkes, 1995b). This process involves
Hm þ Gt oxide− 1 min− 1)
an increase of the surface density of sites susceptible to
acid attack (Watari et al., 1979, 1983; Cornell and S0_60 °C 0 0 0.01037 0.999
S0_180 °C 0 0.01333 0.998
Giovanoli, 1993). At the beginning of dehydroxylation
S0_220 °C 0.414 0.06492 0.966
only a departure of this trend is observed in sample S3. S0_260 °C 1 0.92904 0.997
This behavior will be discussed below. S1_60 °C 2.5 0 0.01386 0.999
S1_180 °C 0 0.02684 0.999
3.6. Congruency of dissolution of Fe and Mn S1_220 °C 0.386 0.09770 0.952
S1_260 °C 1 0.57086 0.991
S2_60 °C 6.7 0 0.01627 0.998
Plots of fraction of Mn versus fraction of Fe S2_180 °C 0 0.02056 0.997
dissolved during dissolution provide an indirect mea- S2_220 °C 0.319 0.07227 0.953
sure of the distribution of Mn within the crystals of S2_260 °C 0.696 0.21034 0.980
goethite and hematite. Dissolution data for Fe and Mn S2_310 °C 1 0.16701 0.998
S3_60 °C 8.3 0 0.02562 0.999
S3_180 °C 0 0.01771 0.974
S3_220 °C 0.286 0.08515 0.969
S3_260 °C 0.641 0.08688 0.998
S3_310 °C 1 0.13368 0.998
(1)
Hm: hematite; Gt.: goethite.

for Mn-substituted goethite and corresponding hematite


are shown in Fig. 8a) and b), respectively.
The data for Mn-substituted goethites (Fig. 8a) is
well described by a line of slope = 1, indicating that
Mn3+ and Fe3+ were uniformly distributed within the
crystal structure of goethite, and that dissolution was
congruent. In contrast, a withdrawal of congruency was
observed for Mn-substituted hematites (Fig. 8b). The
convex χMn : χFe curve for hematite S1_260 °C indicates
that, after the thermal transformation, Mn tends to be
more concentrated towards the surface of the hematite
particles. The opposite trend is observed in S2_310 °C,
which presents a concave curve. In spite of this rather
erratic behavior, it suggests an inhomogeneous Mn
distribution in the hematite phase after thermal treat-
ment, being this effect more significant in samples with
low Mn content. Further studies concerning the crystal
analysis and Mn distribution should be carried out in
order to achieve a better understanding of this process.

3.7. Effect of dissolution temperature–activation energy

The activation energy (Ea) for the dissolution process


was obtained for goethite and hematite phases in samples
S0 (χMn = 0) and S2 (χMn = 0.067), using the Arrhenius
equation, with R2 values ≥ 0.98, (data not shown). The
Fig. 7. Representative dissolution curves for synthetic goethites (a) and
corresponding hematites (b) expressed as mmoles of Fe released per
dissolution rate (k) values were obtained from the slopes
gram oxide in 4 mol dm− 3 HCl at 40 °C vs. time, as affected by Mn of initial nearly linear regions of the dissolution curves
substitution (Mn content, as mol%: S0 = 0; S1 = 2.5; S2 = 6.7; S3 = 8.3). measured at 40, 50 and 60 °C. The calculated Ea value
296 M. Alvarez et al. / Chemical Geology 231 (2006) 288–299

Avrami–Erofejev and the Kabai laws (Cornell and


Schwertmann, 1996). The best description of the kinetic
behavior was found using the Kabai law.
The coefficients obtained fitting the data with the
α
Kabai equation (Kabai, 1973), χFe = 1 − e−(kt) , are listed
in Table 5. Here, χFe is the fraction of Fe dissolved at
time t and α is a constant value. For α N 1 the curve is
sigmoidal in shape, for α b 1 the curve is deceleratory.
The dissolution rate constant, k, was obtained from plots
of ln[ln{1 / (1 − χFe)}] vs. lnt, where ln[ln{1 / (1 − χFe)}]
is the log-transform linearized form of the Kabai
equation.
As can be seen, the dissolution of Mn-substituted
goethites synthesized at 60 °C is well described by the
Kabai equation.
For some but not all samples consisting of goethite
and hematite, v.g. all the samples heated at 220 °C, two
intercepting straight lines were required to achieve a
better fit. The break between the two lines may represent
the point at which most hematite has dissolved so that
continuing a slower dissolution of goethite determines
the slope of the second line.
The same behavior is also detected in samples
S2_180 °C and S3_180 °C, where Mn-goethite is the unique
phase present. This effect is in line with a clear decrease
of the intensity of the low water bending bands, at
∼1650 and ∼1790 cm− 1, with the Mn substitution (Fig.
Fig. 8. Plot of χMn dissolved vs. χFe dissolved in 4 mol dm− 3 HCl for
9). However, we found that the presence of Mn in the
(a) Mn-substituted goethites, and (b) Mn-substituted hematites.
goethite structure retards the complete transformation to
hematite, as shown by XRD, IR and DTA analysis.
for pure goethite (S0_60 °C) is 149.6 kJ mol− 1, while the The good fit of the data found for goethite / hematite
value for pure hematite (S0_260 °C) is 85.6 kJ mol− 1. In mixtures at 260 °C to single lines is difficult to explain.
the case of the substituted goethite, the Ea value en- Although more than one phase is present, there is no
countered for S2_60 °C (90.1 kJ mol− 1) is smaller than assurance of the Mn distribution in each phase, so a
the value calculated for S0_60 °C, this fact may be compensating effect could justify this good fit. The d4
ascribed to the presence of weaker Mn–O bonds com- electron configuration of Mn3+ undergoes a large static
pared to the only Fe–O bonds present in pure goethite Jahn–Teller distortion in octahedral coordination and
(Schwertmann, 1991). this effect may limit the substitution of Mn in hematite.
The comparison between goethite and hematite with The results obtained by Singh et al. (2000) agree with
the same χMn value (S2_60 °C and S2_310 °C) shows Ea those of Cornell and Giovanoli (1987) who showed that
values of 90.1 and 91.0 kJ mol− 1, respectively. The quantities ≤ 5 mol% Mn can be incorporated in the
latter value is in agreement with the Ea values reported hematite structure. Any additional Mn is either adsorbed
by Wells et al. (2001) for hematite synthesized with or precipitated as a separate phase. Since the Jahn–
6.3 mol% Mn substitution (84.1 kJ mol− 1). As can be Teller effect limits the solubility of Mn in hematite, a
seen from the Ea results, Mn substitution produces op- non-uniformity Mn distribution between both phases is
posite effects in the stabilization of goethite and hema- possible.
tite against proton attack. The effect of Mn incorporation in the stability of the
goethite and hematite phases is reflected in the variation of
3.8. Dissolution kinetics the k values obtained from the Kabai law. In all cases, k
shows a significant increase with the total phase transfor-
Dissolution data at 40 °C, for all samples were fitted mation of goethite to hematite; however, this increase is
to several kinetic laws, such as the quadratic, the smaller with the higher Mn content. For example, k derived
M. Alvarez et al. / Chemical Geology 231 (2006) 288–299 297

Table 5
Coefficients for the Kabai equation for samples dissolved at 40 °C
Sample Mol% Mn Hm (1) Linear part n α k (min− 1 × 103) R2
Hm þ Gt
S0_60 °C 0 0 – 7 1.217 1.98 0.994
S0_180 °C 0 – 9 1.282 1.70 0.996
S0_220 °C 0.414 I 11 0.298 0.11 0.997
II 5 0.772 1.61 0.996
S0_260 °C 1 – 9 1.130 81.36 0.991
S1_60 °C 2.5 0 – 7 1.617 3.63 0.999
S1_180 °C 0 – 6 1.189 3.86 0.997
S1_220 °C 0.386 I 8 0.412 7.71 0.989
II 4 1.351 7.80 0.991
S1_260 °C 1 – 10 0.867 54.57 0.999
S2_60 °C 6.7 0 – 9 1.395 2.42 0.997
S2_180 °C 0 I 6 1.139 1.75 0.997
II 3 1.955 2.15 0.978
S2_220 °C 0.319 I 8 0.388 6.42 0.995
II 4 1.064 7.96 0.997
S2_260 °C 0.696 – 12 0.759 23.30 0.997
S2_310 °C 1 – 10 0.926 31.37 0.992
S3_60 °C 8.3 0 – 9 1.016 3.63 0.993
S3_180 °C 0 I 5 0.822 1.53 0.995
II 5 1.722 3.76 0.998
S3_220 °C 0.286 I 7 0.533 8.24 0.993
II 4 1.032 7.30 0.994
S3_260 °C 0.641 – 12 0.784 14.56 0.998
S3_310 °C 1 – 10 1.156 14.37 0.992
(1)
Hm: hematite; Gt: goethite.

from Kabai equation is 1.98×103 min− 1 for pure goethite thermically treated at different temperatures. The degree
(S0_60 °C), and 8.14 × 104 min− 1 for pure hematite of transformation could be evaluated in function of
(S0_260 °C), while for S3_60 °C k value is 3.63×103 and changes in the cell parameters of all samples, and their
1.44 × 104 min− 1 for S3_310 °C. These results are in chemical reactivity in an acidic media.
accordance with k0 values observed in Table 4. Crystals of Mn-substituted goethite retain the acicular
shape after transformation to hematite at 310 °C. The
4. Conclusions temperature of transformation decreases as the Mn con-
tent increases. In the same way, the Mn substitution in
A series of goethites with increasing Mn content were goethite raises the dehydroxylation temperature.
synthesized by forced hydrolysis with base addition, and The cell parameters of both goethite and hematite
decrease with the thermal treatment. However, the be-
havior of the b parameter in Mn-goethite is more erratic.
In agreement with XRD results, IR spectra support
the assumption that Mn retards the dehydroxylation of
goethite solid phase to form hematite.
The results of this study confirm that Mn-substituted
goethite is thermally more stable than non-substituted
goethite. The same effect was observed in Al-goethite.
The acid dissolution of all Mn-goethites is congruent,
but not for all Mn-hematites. The major diversion from
this tendency is observed in hematites with higher Mn
content.
The increasing Mn substitution in goethite raises the
dissolution rate, while the opposite effect was observed
Fig. 9. FTIR spectra from samples S0_180 °C to S3_180 °C in the region of for the corresponding Mn-hematites obtained from heat-
the water bending bands (1600–1900 cm− 1). ing goethite.
298 M. Alvarez et al. / Chemical Geology 231 (2006) 288–299

Mn substitution produces opposite effects in activa- Derie, R., Ghodsi, M., Calvo-Roche, C., 1976. DTA study of the
tion energy for acid dissolution in both goethite and dehydration of synthetic goethite α-FeOOH. J. Therm. Anal. 9,
435–440.
hematite phases. Fey, M.B., Dixon, J.B., 1981. Synthesis and properties of poorly crys-
Dissolution of pure and Mn-substituted goethites and talline hydrated aluminous goethites. Clays Clay Miner. 29, 91–100.
hematites, and samples containing a mixture of both Gasser, U.G., Nüesch, R., Singer, M.J., Jeanroy, E., 1999. Distribution
phases in 4 mol dm− 3 HCl can be well explained by the of manganese in synthetic goethite. Clay Miner. 34, 291–299.
González, G., Sagarzazu, A., Villalba, R., 2000. Study of the mechano-
modified first-order Kabai equation. Data for samples
chemical transformation of goethite to hematite by TEM and XRD.
treated at 180 and 220 °C conform to two straight lines Mater. Res. Bull. 35, 2295–2308.
whereas a single line provided a satisfactory description Goñi-Elizalde, S., García-Clavel, M.E., 1988. Thermal behavior in air
of dissolution of the remaining samples. of iron oxyhydroxides obtained from the method of homogeneous
Further studies such as high resolution TEM should precipitation: Part I. Goethite samples of varying crystallinity.
be carried out in order to elucidate how Mn distribution Thermochim. Acta 124, 359–369.
Goss, C.J., 1987. The kinetics and reaction mechanism of the goethite
in each phase influences the reactivity of these samples. to hematite transformation. Miner. Mag. 51, 437–451.
Jonas, K., Solymar, K., 1970. Preparation, X-ray derivatographic and
Acknowledgements infrared study of aluminium-substituted goethites. Acta Chim.
Acad. Sci. Hung. 66, 383–394.
This work has been partially supported by SGCyT Kabai, J., 1973. Determination of specific activation energies of metal
oxides and metal oxide hidrates by measurement of the rate of
(PGI: 24/005), Universidad Nacional del Sur, Bahía dissolution. Acta Chim. Acad. Sci. Hung. 78, 57–73.
Blanca, Argentina and grant from Universidad de Klissurski, D.G., Bluskov, V.N., 1980. A Mössbauer study of the
Buenos Aires (X168), Argentina. The authors are grate- thermal decomposition of highly disperse α-FeOOH. Mater.
ful to the anonymous reviewers for constructive com- Chem. 5, 67–71.
ments that enhanced the clarity of this manuscript. [DR] Larson, A.C., Von Dreele, R.B., 1994. GSAS. General Structural
Analysis System. Los Alamos Natl. Lab. Rep. LAUR 86, 748.
Lim-Nuñez, R.S.L., 1985. Synthesis and Acid Dissolution of Metal-
References substituted Goethites and Hematites: MSc. Thesis. Department of
Soil and Plant Nutrition, U.W.A., Nedlands, 6009, pp. 104–121.
Alvarez, M., Sileo, E.E., Rueda, E.H., 2005. Effect of Mn(II) on the Lim-Nuñez, R.S.L., Gilkes, R.J., 1987. Acid dissolution of synthetic
transformation of ferrihydrite to goethite. Chem. Geol. 216, 89–97. metal-containing goethites and hematite. Proceedings of the Inter-
Ball, M.C., Taylor, H.F.W., 1961. The dehydration of brucite. Mineral. national Clay Conference, Denver, pp. 197–204.
Mag. 32, 754–766. Lima-de-Faria, J., 1963. Dehydration of goethite and diaspore.
Bernstein, L.R., Waychunas, G.A., 1987. Germanium crystal chem- Z. Kristallogr. 119, 176–203.
istry in hematite and goethite from Apex Mine, Utah, and some Maslen, E.N., Streltsov, V.A., Streltsova, N.R., Ishizawa, N., 1994.
data on germanium in aqueous solution and in stollite. Geochim. Synchrotron X-ray study of the electron density in α-Fe2O3. Acta
Cosmochim. Acta 51, 623–630. Crystallogr., B 50, 435–441.
Blesa, M.A., Morando, P., Regazzoni, A.E., 1994. Chemical Mendelovici, E., Villalba, R., Sagarzazu, A., 1982. Processings of
Dissolution of Metal Oxides. CRC Press, Boca Raton, USA. iron oxides at room temperature: I. Solid state conversion of pure
Cambier, P., 1986. Infrared study of goethite of varying crystallinity α-FeOOH into distinctive α-Fe2O3. Mater. Res. Bull. 17, 241–249.
and particle size: I. Interpretation of OH and lattice vibration Naono, H., Fujiwara, R., 1980. Micropore formation due to thermal
frequencies. Clay Miner. 21, 191–200. decomposition of acicular microcrystals of α-FeOOH. J. Colloid
Cornell, R.M., Giovanoli, R., 1987. Effect of manganese on the Interface Sci. 73 (2), 406–415.
transformation of ferrihydrite into goethite and jacobsite in alkaline Pomiès, M.P., Menu, M., Vignaud, C., 1999. TEM observations of
media. Clays Clay Miner. 35, 11–20. goethite dehydration: application to archaeological samples. J. Eur.
Cornell, R.M., Giovanoli, R., 1993. Acid dissolution of hematites of Ceram. Soc. 19, 1605–1641.
different morphologies. Clay Miner. 28, 223–232. Rietveld, H.M., 1969. A profile refinement method for nuclear and
Cornell, R.M., Schindler, P.W., 1987. Photochemical dissolution of magnetic structures. J. Appl. Crystallogr., B 25, 925–946.
goethite in acid/oxalate dissolution. Clays Clay Miner. 35, Ruan, H.D., Gilkes, R.J., 1995a. Dehydroxylation of aluminous
347–352. goethite: unit cell dimensions, crystal size and surface area. Clays
Cornell, R.M., Schwertmann, U., 1996. The Iron Oxides. Structure, Clay Miner. 43 (2), 196–211.
Properties, Reaction, Occurrence and Uses. VCH, Weinheim Ruan, H.D., Gilkes, R.J., 1995b. Acid dissolution of synthetic
(Germany). 573 pp. aluminous goethite before and after transformation to hematite
Cornell, R.M., Posner, A.M., Quirk, J.P., 1974. Crystal morphology by heating. Clay Miner. 30, 55–65.
and the dissolution of goethite (α-FeOOH). J. Inorg. Nucl. Chem. Ruan, H.D., Frost, R.L., Kloprogge, J.T., 2001. The behavior of
36, 1937–1946. hydroxyl units of synthetic goethite and its dehydroxylated product
Cornell, R.M., Posner, A.M., Quirk, J.P., 1975. The complete hematite. Spectrochim. Acta Part A 57, 2575–2586.
dissolution of goethite. J. Appl. Chem. Biotechnol. 25, 701–706. Schwertmann, U., 1984a. The influence of aluminium on iron oxides:
Cornell, R.M., Posner, A.M., Quirk, J.P., 1976. Kinetics and IX. Dissolution of Al-goethite in 6 M HCl. Clay Miner. 19, 9–19.
mechanisms of the acid dissolution of goethite (α-FeOOH). Schwertmann, U., 1984b. The double dehydroxylation peak of
J. Inorg. Nucl. Chem. 38, 563–567. goethite. Thermochim. Acta 78, 39–46.
M. Alvarez et al. / Chemical Geology 231 (2006) 288–299 299

Schwertmann, U., 1991. Solubility and dissolution of iron oxides. Surana, V.S., Warren, I.H., 1969. The leaching of goethite. Trans. Inst.
Plant Soil 130, 1–25. Min. Metall. 80, C152–C155.
Schwertmann, U., Cornell, R.M., 1991. Iron oxides in the laboratory. Szytula, A., Burewicz, A., Dimitrijevic, Z., Krasnicki, S., Rzany, H.,
Preparation and Characterization. VCH, Weinheim. 137 pp. Todorovic, J., Wanic, A., Wolski, W., 1968. Neutron diffraction
Schwertmann, U., Lathman, M., 1986. Properties of iron oxides in studies of α-FeOOH. Phys. Stat. Solidi 26, 429–434.
some new Caledonian oxisols. Geoderma 39, 105–123. Thompson, P., Cox, D.E., Hastings, J.B., 1987. Rietveld refinement of
Sidhu, P.S., Gilkes, R.J., Cornell, R.M., Posner, A.M., Quirk, J.P., Debye–Scherrer synchrotron X-ray data from Al2O3. J. Appl.
1981. Dissolution of iron oxides and oxyhydroxides in hydro- Crystallogr. 20, 79–83.
chloric and perchloric acids. Clays Clay Miner. 29, 269–276. Trolard, F., Bourrie, G., Jeanroy, E., Herbillon, A.J., Martin, H., 1995.
Sileo, E.E., Alvarez, M., Rueda, E.H., 2001. Structural studies on the Trace metals in natural iron oxides from laterites: a study using
manganese for iron substitution in the synthetic goethite–jacobsite selective kinetic extraction. Geochim. Cosmochim. Acta 59,
system. Int. J. Inorg. Mater. 3, 271–279. 1285–1297.
Sileo, E.E., Ramos, A.Y., Magaz, G.E., Blesa, M.A., 2004. Long-range Watari, F., van Landuyt, J., Delavignette, P., Amelinckx, S., 1979.
vs. short-range ordering in synthetic Cr-substituted goethites. Electron microscopic study of dehydration transformations: I.
Geochim. Cosmochim. Acta 68 (14), 3053–3063. Twin formation and mosaic structure in hematite derived from
Singh, B., Gilkes, R.J., 1992. Properties and distribution of iron oxides goethite. J. Solid State Chem. 29, 137–150.
and their association with minor elements in the soils of south- Watari, F., Delavignette, P., van Landuyt, J., Amelinckx, S., 1983.
western Australia. J. Soil Sci. 43, 77–98. Electron microscopic study of dehydration transformations: III. High
Singh, B., Sherman, D.M., Gilkes, R.J., Wells, M., Mosselmans, resolution observation of the reaction process FeOOH → Fe2O3.
J.F.W., 2000. Structural chemistry of Fe, Mn and Ni in synthetic J. Solid State Chem. 48, 49–64.
hematites as determined by extended X-ray absorption fine Wells, M.A., Gilkes, R.J., Fitzpatrick, R.W., 2001. Properties and acid
structure spectroscopy. Clays Clay Miner. 48, 521–527. dissolution of metal-substituted hematites. Clays Clay Miner. 49
Stephens, P.W., 1999. Phenomenological model of anisotropic peak (1), 60–72.
broadening in powder diffraction. J. Appl. Crystallogr. 32, Wolska, E., Szajda, W., 1985. Structural and spectroscopic character-
281–289. istics of synthetic hydrohaematite. J. Mater. Sci. 20, 4407–4412.

You might also like