You are on page 1of 9

Applied Clay Science xxx (xxxx) xxx–xxx

Contents lists available at ScienceDirect

Applied Clay Science


journal homepage: www.elsevier.com/locate/clay

Research paper

Identifying the differences between clays used in the brick industry by


various methods: Iron extraction and NMR spectroscopy

Julie Peynea,b, Ameni Gharzounia, Isabel Sobradosc, Sylvie Rossignola,
a
Univ. Limoges, IRCER, UMR 7315, F-87000 Limoges, France
b
Bouyer Leroux, L'établère 44210, La Séguinière, France
c
Instituto de Ciencia de Materiales de Madrid, Consejo Superior de Investigaciones Científicas (CSIC), C/Sor Juana Inés de la Cruz, 3, 28049 Madrid, Spain

A R T I C LE I N FO A B S T R A C T

Keywords: This study focused on identifying the differences in a clay quarry that is used in brick production to understand
Clay minerals the variation between samples based on two kinds of clays: the first kind was suitable for brick production, but
Iron the second one did not show good production efficiency. Various samples were characterized by physical and
NMR spectroscopy chemical methods. In addition, free iron extraction by a CBD method and calcination at various temperatures
Calcination
were performed. The clay samples were then structurally characterized by XRD, FTIR and 27Al NMR spectro-
Bricks clay
scopy and thermally characterized using dilatometry. Finally, these results were correlated to the SEM in-
vestigation. These results revealed variations in the Si/Al molar ratio (1.4 to 3.3) due to the mineralogy variation
within the quarry. Indeed, the non-secondary phase and clay phase contents differed according to the location in
the quarry. NMR analysis also revealed that some iron atoms were located in a kaolinite structure, whereas other
iron phases, in goethite form, interacted with clay minerals only on the surface of, e.g., kaolinite and muscovite
and were not in their structures. From the thermal characterization, the influence of this structural iron in the
kaolinite was also evidenced by an increase in the total shrinkage. In conclusion, this work determined the
structural differences and their impact on brick production.

1. Introduction preparation (Jimenez-Quero et al., 2017). Even if it were possible to


separately determine the contribution of each mineral to the properties
Clay minerals are characterized by several physical and chemical of the mixture, the characterization techniques used in the industry
properties, such as their swelling behavior, cation exchange capacity (mainly the particle size distribution, chemical analysis and miner-
(CEC), cation adsorption, and high surface area (Bergaya and Lagaly, alogy) would not allow the suitability of the clay for brick production to
2013) that cause these raw materials to be used strategically in nu- be determined. The most advanced techniques, such as 27Al NMR and
merous industrial applications, such as brick production (Gualtieri SEM microscopy, are expensive and thus not yet in use. However,
et al., 2015). several previous reports demonstrated that 27Al NMR spectroscopy is a
Conventionally, bricks are made from a mixture of numerous clays very powerful tool for deeply investigating the crystal chemistry of
extracted from different quarries (Kornmann, 2009). The final proper- layered silicates such as kaolinite, montmorillonite, and muscovite
ties of the bricks (e.g., presence of failures, color, durability, compres- (Thompson, 1984, Woessner, 1989, Takahashi et al., 2008). For ex-
sive strength) clearly depend on the raw materials used. Nonetheless, ample, it is possible to evaluate the evolution of the 6-, 5- and 4-co-
the non-homogeneous mineralogy content of the brick clay mixture ordinate aluminum sites with temperature during the dehydroxylation
creates difficulties in understanding the impact of each mineral from reaction of kaolinite (Benharrats et al., 2003; Wanabe et al., 1987), thus
each quarry. However, it is well known that an increase in the mon- revealing the conversion of the 6-coordinate into 5- and 4-coordinate
tmorillonite content leads to technical issues during the drying step due aluminum. The structures of the clays considered for this application
to increases in the shrinkage and capillarity retention (Sigg, 1991). In are not often described in the literature. However, Mackenzie et al.
addition, the final color of the brick is clearly due to the initial iron (1987) studied the thermal transformations of muscovite and used 27Al
content (Li et al., 2016). The presence of kaolinite also increases the and 29Si NMR to demonstrate that iron impurities lead to an increase in
plasticity of the brick mixture, thereby improving the mixture recrystallization and favor the mullite to corundum conversion. In


Corresponding author.
E-mail address: Sylvie.rossignol@unilim.fr (S. Rossignol).

https://doi.org/10.1016/j.clay.2018.02.037
Received 12 October 2017; Received in revised form 23 February 2018; Accepted 25 February 2018
0169-1317/ © 2018 Elsevier B.V. All rights reserved.

Please cite this article as: Peyne, J., Applied Clay Science (2018), https://doi.org/10.1016/j.clay.2018.02.037
J. Peyne et al. Applied Clay Science xxx (xxxx) xxx–xxx

addition, Wei et al., 2012 highlighted the impact of goethite in a kao- identified by XRD with a BRUKER AXS D8 Advance powder dif-
linite-goethite mixture or in a goethite interaction with kaolinite, re- fractometer using CuKα radiation (λα = 0.154186 nm). The in-
vealing a 6-coordinate aluminum contribution at a lower chemical shift vestigated 2theta range used was between 2.5 and 65°, with a step size
for the goethite-kaolinite mixture than for the goethite interaction with of 0.012° (2theta) and an acquisition time of 2 s per step. JCPDS files
kaolinite (3.04 and 5.76 ppm, respectively). It is also possible to detect (Joint Committee Powder Diffraction Standard) were used for the
the kaolinite default in the structure and, more precisely, the iron in the principal phase identification.
kaolinite structure, as described by Schroeder and Pruett, 1996. 27Al High-resolution MAS-NMR experiments were performed at room
NMR is thus an appropriate structural characterization method for temperature using a Bruker AVANCE-400 spectrometer operating at
studying the impact of impurities on clay minerals. It is clear that the 104.26 MHz (27Al signal) on powder samples, which were spun at
iron oxide and (hydr)oxide contents can influence the clay structure. 10 KHz. The number of scans was 400(Autef et al., 2013). In order to
Differences within a clay quarry could thus occur (Guzlena et al., 2017) record the central transitions of aluminum (27Al) (I = 5/2), a π/8
and lead to technical issues after numerous steps of brick production. (1.5 μs) pulse was applied to it and 1 MHz filter was used. A recycle
Thus, this study aims to understanding the differences in a clay quarry delay of 5 s was chosen to minimize saturation effects. Chemical shift
and their influence on the quality of brick production. To achieve this values were given with reference to an external aqueous AlCl3 solution.
goal, two lots of clay with different impacts on brick production were The relative areas of the components were determined using DMFIT
extracted in northwest France. Chemical treatments, such as a free iron program.
extraction and thermal treatment (with various calcinations at two Diffuse reflectance spectra of the clay minerals (in powder form)
temperatures: 350 and 700 °C), were then performed to demonstrate were obtained between 300 and 780 nm with a CARY 500 spectro-
the influence of the iron and the thermal behavior of each clay. Next, photometer (cf. Supplementary files). A remission function (also called
the physical and chemical features and structural information were the Kubelka Munk function, Kubelka and Munk, 1931; Barron and
obtained by X-ray powder diffraction, FTIR and 27Al NMR spectroscopy. Torrent, 1986) was applied to the experimental data according to the
(1 − R)2
These results were then correlated to the thermal behavior of the clays literature, with f (R) = 2R , where R is the reflexion. Noise reduction
before and after the chemical and thermal treatments. was then performed by fitting each Kubelka Munk spectrum with an
Adjacent Averaging (5 points) method. Finally, the second derivate was
2. Experimental calculated using Origin V8.
Dilatometric measurements were made under air by means of a
2.1. Raw materials and samples preparation contact horizontal dilatometer (NETZSCH DIL 402C) on clay samples
pressed with a cylindrical geometry (H = 10 mm; Ø = 6.5 mm). Two
Numerous samples (18) from one quarry that is located in northwest alumina (Al2O3) holders were placed at the surface end contacts of the
France and used in the brick industry were selected for this study. Two samples to prevent high-temperature diffusion between the samples and
different lots of samples (A and B) with two distinct sets of visual the alumina rod. A calibration cycle (with a standard sample of Al2O3)
characteristics, such as color and grain size, were observed. They were was performed and registered. The calibration data were then sub-
also selected according to their influence on brick production, i.e., on tracted from the data collected for each sample to eliminate the con-
the scrap rate after the sintering of bricks. Indeed, A samples were tribution from the device. The thermal cycle used consisted of heating
suitable for brick production, but B samples were not suitable because at 10 °C/min to 1100 °C and cooling to 25 °C at a similar cooling rate.
they led to a significant increase in the scrap rates after the sintering
step. The sampling was performed based on the height of the rock face.1
The preparation of these samples was divided into two steps: (i) drying 3. Results and discussion
for 24 h at 40 °C and (ii) coarse grinding (< 1 mm). The red/yellow/
brown color of these samples clearly demonstrates the presence of iron 3.1. Physical and chemical characterization
oxohydroxide (Cornell and Schwertmann, 2002). Using the Mehra and
Jackson procedure (Mehra and Jackson, 1960), free iron oxide (organic The physical and chemical features of two sample sets (y = 1, 2, 3,
complexes, amorphous and highly crystallized forms) can be totally 4) for the Ay-25 and By-25 lots are presented in Table 1. No significant
dissolved. The main reactants used are a Na-citrate solution (0.3 M) and variation is observed in the weight percentage of the > 40 μm granu-
a NaHCO3 solution (1 M), with sodium dithionite (Na2S2O4). A suitable lometry range (31–45 and 36–46 wt% for By and Ay, respectively).
amount of the clay sample (4 g) was mixed with 40 mL of the Na citrate Nonetheless, an increase in the total fraction of the 40–80 μm range can
solution and 5 mL of 1 M NaHCO3 in a water bath (80 °C). Then, 1 g of be observed for the By lot (8–15 and 5–6 wt% for Ay and By, respec-
Na2S2O4 was added, mixed continuously for 1 min and then mixed tively). The specific surface area (SBET) and cation exchange capacity
occasionally for 15 min. Then, 10 mL of a saturated NaCl solution was (CEC) values depend on the sampling lot, and they vary with the same
added and centrifuged for 5 min (at 2500 rpm/min). Finally, the su- behavior. Indeed, higher SBET and CEC values are obtained for Ay than
pernatant was decanted into a flask, and the residual solid was kept and By (27 m2/g − 11.6 cmol(+)/kg and 17 m2/g − 6.4 cmol(+)/kg, re-
dried at 40 °C for 24 h. To characterize the temperature behavior, var- spectively). The wettability values (measurement previously explained
ious calcinations in an electrical furnace for bulk samples before and in Peyne et al., 2017) also depend on the sampling lot, with higher
after the iron extraction were performed at 350 and 700 °C (T), which values for By-25 (162–235 and 129–146 μL/g for By-25 and Ay-25, re-
corresponded to the temperatures after iron (hydr)oxide decomposition spectively). The chemical compositions are consistent and exhibit a
and clay minerals dihydroxylation, respectively (Ruan et al., 2002; significant change, with higher Si/Al and Si/Fe molar ratios for the Ay
Lopez et al., 2011). The following nomenclature is used: A/By-C-T, samples. Fig. 1 shows pictures of the samples before and after the
where y specifies the number of the sample, C is the performed che- chemical treatment and calcination. The yellow color of the A1–25
mical treatment and T is the calcination temperature. sample is typical for a rich goethite soil (Cornell and Schwertmann,
2002). The calcination at 350 and 700 °C (A1–350 and A1–700) leads to a
change from yellow to a red color, assigned to the goethite to hematite
2.2. Characterization techniques
transformation (Carlos Duarte Cavalcante et al., 2017). The free iron
extraction induces a change in color from yellow/orange to gray/white,
The mineral phases of the samples (either calcined or not) were
in agreement with the protocol, which shows that a high amount of
goethite content was removed. However, calcination at 350 °C for this
1
Bouyer Leroux, L'établère, 49,280 La Séguinière, France. sample does not lead to a significant change, with only a slightly darker

2
J. Peyne et al. Applied Clay Science xxx (xxxx) xxx–xxx

Table 1
Physical and chemical properties of various samples.

Samples Particle size distribution Physical and chemical features Chemical composition

[ > 40 μm] (%) ± 1 [40–80 μm] wt% ± 1 SBET (m2/g) ± 1 CEC (cmol(+)/kg) ± 0.5 W (μL/g) ± 5 Si/Al molar ratio Si/Fe molar ratio

B1–25 45 15 17 6.4 162 1.5 8.1


B2–25 31 8 20 6.1 235 1.4 8.3
A3–25 46 5 27 11.6 146 2.1 12.8
A4–25 35 6 34 8.6 129 3.3 16.1

(A) (B)
A1-T A1-C-T B1-T B1-C-T

(a) T=25 °C

(b) T=350 °C

(c) T=700 °C

Fig. 1. Pictures of (A) A1-T and A1-C-T; (B) B1-T and B1-C-T samples for various temperatures T = (a) 25, (b) 350 and (c) 700 °C.

gray color (A1-C-350). It is only after calcination at 700 °C that a red 2001; Vaculikova et al., 2011), and one at 3619 cm−1, which corre-
coloration is observed for the sample (A1-C-700). Similar changes in color sponds to inner the Al2-OH stretching (Vaculikova et al., 2011) of
were detected for the B1-T and B1-C-T samples. However, the A1–700 kaolinite minerals (Parker, 1969). The band at 1642 cm−1 is directly
sample looks darker than the B1–700 sample. In addition, the B1-C-T linked to the OeH vibration band from the water and is mainly related
samples seem darker than the A1-C-T samples. All these differences in to the preparation of the KBr pellets. A broad band at 1172 cm−1 is
color between the lots could be due to either a different type of iron observed and is due to the SieOeSi from quartz (Ravinska et al., 2010).
(hydr)oxide or different local coordination sphere for the iron (Her- The other contributions at 1106, 1031, and 1004 cm−1 are related to
billon et al., 1976). For the following results description, the number of kaolinite and to the SieOeSi stretching vibrations (Oikonolopoulis
the sampling will be deleted from the nomenclature to simplify the text. et al., 2015). Additionally, the bending vibrations of the inner Al2-OH
UV spectroscopy was performed to characterize the type of (hydr) and the SieOeAl vibration band from kaolinite are observed at 912 and
oxides in the samples after iron extraction (C) and after thermal treat- 533 cm−1, respectively (Vaculikova et al., 2011). The SieO vibration
ment. The data obtained are presented in the supplementary file and bands from quartz minerals can be seen at 796, 778 and 692 cm−1
show that increasing the calcination temperature induces a decrease in (Ravinska et al., 2010). Similar contributions are observed for the AC/25
goethite and an increase in hematite. Moreover, the amount of goethite sample (Fig. 2a). An additional contribution is observed at 1402 cm−1,
is slightly lower for the A350 than for the B350 samples, suggesting a which can be attributed to the sodium carbonate vibrations (Huang and
higher transformation of goethite to hematite at this temperature, in Kerr, 1960) arising from the iron extraction procedure. A weak con-
agreement with the darker color relative to the higher amount of he- tribution can also be seen after the iron extraction at 933 cm−1, which
matite content (Table 1). After the free iron extraction for all samples may be correlated to the inner surface Al2-OH. With the exception of
(AC,T or BC,T), the changes differ because goethite is not detected, some contributions, the B25 raw clay sample shows a similar FTIR
consistent with the chemical extraction (see Figs. 1 and 2 in the sup- spectrum (Fig. 2b). However, the two absorption bands at 3669 and
plementary file). A hematite contribution only appears at 700 °C, which 3642 cm−1 due to kaolinite are more defined, which clearly means the
seems to be linked to clay mineral dehydroxylation (such as mica, kaolinite is highly crystalline (Felhi et al., 2008). In addition, the quartz
kaolinite). However, this characterization technique does not allow the doublet at 792 and 778 cm−1 is less visible, which suggests a lower
determination of the iron coordination sphere in this clay mineral. quartz content. The free iron extraction does not lead to any change in
the FTIR spectrum (Fig. 2b).
The X-ray powder diffraction patterns of the clay samples at various
3.2. Structural characterization calcination temperatures and before/after the free iron extraction are
plotted in Fig. 3 (a, b). The A25 sample (Fig. 3a) is typical of a clay
To evaluate the influence of the iron extraction on the clay minerals, mixture composed of quartz and kaolinite but also muscovite and
FTIR spectroscopy was performed. The FTIR spectra of the samples feldspar (albite and anorthite). The goethite content observed in the
before and after the free iron extraction are shown in Fig. 2 (a, b). The UV–visible spectra (data shown in supplementary file) is not detected
A25 sample (Fig. 2a) exhibits three contributions at 3692, 3669 and due to its small amount (in comparison with quartz) and to the poor
3642 cm−1, which can be attributed to the OH stretching of inner- crystallinity of this type of mineral, usually observed in soils
surface hydroxyl groups (Madejova, 2003; Madejova and Komadel,

3
J. Peyne et al. Applied Clay Science xxx (xxxx) xxx–xxx

(a) 80 and 40 ppm and another in the range of 20 to −20 ppm, which are
attributed to four- and six-coordinate aluminum, respectively (Bardy
et al., 2007). The 27Al NMR spectra of the A25 clay sample (Fig. 4a)
shows two contributions at 68 and 56 ppm, due to four-coordinate
aluminum (Al1IV IV
68ppm and Al256ppm) and two contributions related to six-
coordinate aluminum (Al11ppm and Al2VI
VI
−6ppm) (Sanchez-Munoz et al.,
2013). The signal at 68 ppm may be related to Al in Q3(3Si) sites,
whereas the signal at 56 ppm may be related to Al in Q4(4Si) sites, as
suggested by Takahashi et al. (2008), in the structural framework of
montmorillonite.
A similar spectrum is obtained after calcination at 350 °C (A350 -
Fig. 4a). However, the intensities of the Al1IV IV
68ppm and Al256ppm peaks
decrease slightly. Calcination at 700 °C (A700 – Fig. 4a) leads to the
appearance of a five-coordinate aluminum contribution at 30 ppm
(AlV), linked to the dehydroxylation of clay-type minerals such as
kaolinite and muscovite (Massiot et al., 1995; Carroll et al., 2005). In
addition, it appears that the Al2IV 56 ppm contribution increases but that
the Al1IV68 ppm contribution is not detected. Iron extraction (AC-25–
Fig. 4a) leads to similar spectra to A25. However, an important increase
4000 3500 3000 1500 1000 500 in the Al1IV IV
68 ppm and Al256 ppm contributions at the expense of the six-
coordinate aluminum contributions (Al1VI VI
1 ppm and Al2-6 ppm) is observed
Wavenumber (cm-1) after calcination at 350 °C (AC-350 – Fig. 4a). Comparable behavior is
observed and is even accentuated after calcination at 700 °C (AC-700–
(b) Fig. 4a). In addition, a new contribution at 40 ppm is observed, which
could be attributed to another four-coordinate aluminum (Al3IV). It also
seems that the appearance of the five-coordinate aluminum contribu-
tion is more pronounced for this sample. The Al1VI contribution cannot
be detected after calcination at 700 °C and free iron extraction. The
major contributions observed for the B25 clay mineral (Fig. 4b) are due
to six-coordinate aluminum environments (at 1 and −6 ppm, Al1VI and
Al2VI, respectively), whereas those at 68 and 56 ppm for four-coordinate
aluminum are minor (Al1IV IV
68ppm and Al256ppm). Calcination at 350 °C
leads to a weak increase in these contributions (B350 - Fig. 4b). The
main differences between the A700 and B700 samples occurs in the in-
tense bands of the 5-coordinate aluminum (AlV) at 30 ppm and of the 4-
coordinate aluminum (Al1IV IV
68ppm and Al256ppm), to the detriment of the
contributions of the 6-coordinate aluminum (Al1VI VI
1ppm and Al2-6ppm).
Equivalent behaviors are observed for the BC-25 and BC-350 samples
(Fig. 4b). However, B700 and BC-700 exhibit important differences, par-
ticularly an important diminution of the AlV contribution and a higher
transformation rate of the 6-coordinate aluminum to the 4-coordinate
aluminum (Al1VI VI IV IV
1ppm and Al2-6ppm in Al168ppm and Al256ppm). In addition,
4000 3500 3000 1500 1000 500
the contribution of the newly formed 4-coordinate aluminum is more
-1
Wavenumber (cm ) pronounced. A more detailed description, with the FWMH and areas of
the various contributions for each NMR spectrum, is provided in
Fig. 2. FTIR spectra of the (a) e A25 and AC-25 (b) e B25 and BC-25 samples. Table 2. It seems that at all temperatures studied, the AlVI -6ppm con-
tribution areas are higher for AT and AT/iron. This can be attributed to a
(Vodyanitskii, 2010). The free iron extraction (AC-25) is characterized local environment of kaolinite or of Al-substituted goethite, which may
by similar reflections. The 350 °C calcination (A350 and AC-350) does not be observed in soil (Kim et al., 2015). In addition, the Al1IV 68ppm and

induce any change in the X ray powder patterns (Fig. 5a). However, the Al2IV
56ppm contributions are significantly higher. This can be attributed to

700 °C calcination (A700 and AC-700) leads to disappearance of the the feldspar content, in agreement with the X-ray diffraction patterns
kaolinite reflections, in agreement with kaolinite dehydroxylation (Fig. 3). For all samples, it appears that the iron extraction enhances the
(Lopez et al., 2011) and the additional appearance of a dome between transformation of the six coordinate aluminum to the 4-coordinate
20 and 30°(2θ). The characterization of the BT and BC-T samples is si- aluminum (from 26.2 to 65.5% and from 6.9 to 76.8% for A700 and AC-
milar to that of the AT and AC-T samples, respectively (Fig. 5b), although 700, respectively). The new 4-coordinate aluminum that appears after

the feldspar reflections are less apparent. Moreover, the kaolinite and calcination at 700 °C may be due to secondary or intermediate phases
muscovite reflections of BT and BC-T are more intense than they are for such as Al-O-Ca (Pardal et al., 2012; Wel-Lin Shao, 1989). The calcium
AT and AC-T. These qualitative data from X-ray diffraction do not pro- could come from calcite decomposition, which was not detected by X-
vide a final conclusion about the iron coordination sphere in the clay ray powder diffraction due its low amount (Stanmore and Gilot, 2005;
minerals. Pardal et al., 2012; Wel-Lin Shao, 1989), and could then react with
To better characterize the iron coordination sphere and especially other compounds. In the same way, the Al2IV contribution, mainly from
the iron position in the clay structure, a spectroscopic study was per- muscovite, is greater for A350 than it is for B350. An increase from 8.4 to
formed using 27Al NMR. The 27Al NMR spectra of the clay minerals after 23.4% for A350 and AC-350 is observed and may indicate that the free
calcination and after free iron extraction are presented in Fig. 4 (a, b). iron inhibits this local environment for the A samples, but this is not
All the clay samples are characterized by a broad contribution between observed for the B samples. For the previous samples at 700 °C, the
decrease in the Al1VI 1 ppm contribution and the appearance of the Al
V

4
J. Peyne et al. Applied Clay Science xxx (xxxx) xxx–xxx

(a) (b)

Q
Q Q
Q
I Q Q I I
I I I I I I I I I I I I Q
I A I I I I IQ
I Kaol
I I
I Kaol I A
AC-700 BC-700

A700 B700

AC-350 BC-350

A350 B350

AC-25 BC-25

A25 B25

3 5 10 15 20 25 30 35 40 3 5 10 15 20 25 30 35 40
2θ (°) 2θ (°)

Fig. 3. XRPD patterns of (a) AT, AC-T and (b) BT, BC-T samples (I: Illite, Kaol: Kaolinite, Q: Quartz, A: Albite) (PDF files of I: 00-003-0849, Q: 01-078-1252, Kaol: 00-003-0052, A: 04-014-
0490).

contribution are strong and are linked to kaolinite dehydroxylation. In 3.3. Thermal behavior
addition, this observation is related to the increase in the AlIV con-
tributions, which are consistent with iron that could be linked to the To observe the influence of the iron chemical speciation in the two
kaolinite sheet. clays, the A25, AC-25, B25 and BC-25 samples were characterized by

(A) (B)

Al2(IV)
Al2(IV) (IV) Al3(IV)
Al1
Al3(IV)
Al1(IV)
Al1 (VI)
Al(V) Al1(VI)
Al(V) Al2(VI) Al2(VI)
AC-700 BC-700

A700
B700

AC-350
BC-350

A350
B350

AC-25
BC-25

A25
B25

100 80 60 40 20 0 -20 -40 100 80 60 40 20 0 -20 -40

(ppm) (ppm)

27
Fig. 4. Al NMR spectra of the samples (A) AC-T, AT and (B) BC and BC-T for T = 25, 350 and 700 °C.

5
J. Peyne et al. Applied Clay Science xxx (xxxx) xxx–xxx

Table 2
27
FWHM and area of the various contributions from Al MAS-NMR spectra.

Sample Al1 IV
68 ppm Al2IV 56 ppm Al3IV 40 ppm AlV 30 ppm Al1VI 1 ppm Al2VI −6 ppm

FWHM (ppm) Area (%) FWHM (ppm) Area (%) FWHM (ppm) Area (%) FWHM (ppm) Area (%) FWHM (ppm) Area (%) FWHM (ppm) Area (%)

A25 14 7.2 10 12.4 13 72.7 12 7.8


AC-25 12 5.8 9 11.8 12 68.6 12 13.8
A350 14 5.1 9 8.4 12.5 80.7 12 5.8
AC-350 12 10.1 11 23.4 12 58.2 15 8.3
A700 10 65.5 30 6,9 12 26.2 5.2 1.3
AC-700 14 76.8 14 6.1 33 10.3 10 6.9
B25 18 5.6 12 1.1 13 88.0 12 5.2
BC-25 14 5.3 18 3.9 13 86.0 12 4.6
B350 16 7.2 12 2.2 13 88.8 12 1.8
BC-350 12 5.7 16 3.6 13 89.8 12 1.2
B700 12 6.0 12 11.5 45 37.2 14 35.4 12 1.1
BC-700 14 19.1 14 36.3 12 9.8 30 28.7 14 6

(a) (b)
4 4

2 2
Dilatation (%)

0 Dilatation (%) 0
30 200 400 600 800 1000 1100 30 200 400 600 800 1000 1100

-2 Temperature (°C) -2 Temperature (°C)

-4 -4

-6 -6

-8 -8

Fig. 5. Dilatometry measurements with heating of samples before and after the iron removal for (a) e A25, AC,25 and (b) e B25, BC,25.

dilatometry (Fig. 5 a, b). A ratio higher than 1 is characteristic of an AlX contribution that is less
For all the samples, two shrinkages are observed, the first at ap- important after the iron extraction than before the extraction. A ratio
proximately 900 °C due to mullite formation and the other at 1050 °C smaller than 1 indicates an AlX contribution that is greater after the iron
due to the free silica produced by the thermal decomposition of the clay extraction than before. The contribution ratios of the Al2VI for A25/AC/
minerals, which delays the viscous flow (Castelein et al., 2000). The A25 25 and A350/AC/350 (▼- Fig. 6A–a) are constant and < 1. However, an
clay (Fig. 5a) is characterized by a quasi-null total shrinkage, and the increase in this ratio is observed between 25 and 350 °C for the BC/BC-T
B25 clay (Fig. 5b) shrinkage is more pronounced (up to 2%). After the contributions (Fig. 6A–b). The decrease in the Al2VI contribution of the
iron extraction, the dimensional variation of the AC-25 sample is similar BC,350 sample could be linked either to the some aluminum that is in-
to the A25 sample, while the BC-25 sample (Fig. 5b) is characterized by itially in a highly distorted environment in the sample (Newman et al.,
an increase in the total shrinkage (6%). This reveals that the free iron in 1994) or to the presence of some gibbsite minerals in the initial sample
the AC-25 sample does not perturb the sintering phenomena of the clay (Ashbroook et al., 2000). In addition, the increase in the Al2VI con-
phases (Fig. 5a). However, this chemical treatment leads to the ap- tribution after the iron extraction for the AC/T sample (ratio < 1) may
pearance of these two shrinkages at slightly lower temperatures for BC- mean that the free iron was more extractable. This is consistent with the
25. It seems that during the dehydroxylation of the kaolinite, the pre- darker color of the A350 clay compared to B350, where the free iron was
sence of iron in the clay structure (observed by the X-ray mapping, see more susceptible to transformation into hematite at this temperature.
Supplementary files) leads to the formation of some iron oxide, which Indeed, the lighter color could be due to some iron phases. For all
favors sintering (Soro, 2003). This results in a lower sintering tem- samples, the ratio of the Al1VI contributions (▲- Fig. 6A), attributed to
perature, consequently increasing the total shrinkage. Finally, the the 6-coordinate aluminum of kaolinite, remains constant between 25
presence of some structural iron inside the kaolinite structure can be and 350 °C and increases after the 700 °C calcination. At 700 °C, this
linked to the increase in the scrap rate during brick production. Indeed, ratio is slightly higher for the B700/BC,700 ratio than for A700/AC,700 (5.9
the increase of the shrinkage at high temperature may induce higher and 3.8, respectively), which suggests a more significant dehydroxyla-
stress, leading to scrap in the final shape of the bricks. tion process for the B clay sample after the iron extraction. The Al1VI
transformation into Al1,2IV aluminum is more favorable in case of the B
sample, which seems to reveal an interaction between the free iron and
4. Discussion
the kaolinite content. The ratio of the AlV contributions at 700 °C ( -
Fig. 6B – a,b), higher for B700/BC-700 than for A700/AC-700 (1.36 and
The 27Al NMR results reported above showed some differences,
0.65, respectively), is characteristic of more rapid kinetics of the kao-
depending on the temperature, for the 6-coordinate aluminum (Al1VI
linite dehydroxylation (from a 6-coordinate Al to a 4-coordinate alu-
and Al2VI). Based on this result, the ratios between the contributions
minum) after the iron extraction for the B sample. This is consistent
before and after the iron extraction were calculated for the 4-, 5- and 6-
with the previous conclusion. The variation with temperature of the
coordinate aluminum for AT/AC-T and BT/BC-T samples (Fig. 6A,B – a,b).

6
J. Peyne et al. Applied Clay Science xxx (xxxx) xxx–xxx

(A)
(a) (b)

7 7

6 6

5 5

4 4
AlVI(%) (AT)

AlVI(%) (AC/T)

AlVI(%) (BT)

AlVI(%) (BC/T)
3 3

2 2

1 1

0 0
0 100 200 300 400 500 600 700 800 0 100 200 300 400 500 600 700 800

Temperature (°C) Temperature (°C)

(B)

7 7

6 6

5 5
AlIV-V(%) (AC/T)
AlIV-V(%) (AT)

4 4
AlIV-V(%) (BC/T)
AlIV-V(%) (BT)

3 3

2 2

1 1

0 0
0 100 200 300 400 500 600 700 800 0 100 200 300 400 500 600 700 800

Temperature (°C) Temperature (°C)

Fig. 6. Ratio of the (A) AlVI contribution and (B) AlIV and AlV, before and after the iron oxide removal for (a) AT, AC,T and (b) BT, BC,T with ▼ Al2VI; ▲Al1VI; ●Al2IV; ■ Al1IV and AlV.

ratio of the Al1IV contributions (●- Fig. 6B) differs in the two samples. in interaction with some iron phases, which may decrease the efficiency
As a function of temperature, the ratio for the B samples remains con- of the kaolinite dehydroxylation.
stant at approximately 0.3 (Fig. 6Bb), but for the A samples, a small
decrease is observed from 1.12 to 0.35 for 25 and 350 °C, respectively 5. Conclusions and theoretical model
(Fig. 6Ba). This shows that iron extraction favors the transformation of
6-coordinate aluminum to 4- and 5-coordinate aluminum. The change Based on all these characterization results, it is possible to establish
in the ratio of the Al2IV contributions with temperature is different for a theoretical model of the two clays (A and B) from the same quarry
the A and B samples (Fig. 6B). Between 25 and 350 °C, the ratio de- (Fig. 7). Calcination at 350 and 700 °C leads to different colors for the
creases from 1.3 to 0.5, respectively, for the A samples, but for the B two lots of clays. However, UV spectroscopy only confirms the presence
samples, a slight increase from 1.02 to 1.31 occurs at 25 and 350 °C, of goethite and its transformation into hematite after calcination for
respectively. At T = 700 °C, only the contributions for B700 and BC-700 both clays. This technique did not yield any conclusions about the co-
are observed, which may be consistent with an easier kaolinite dehy- ordination sphere of the iron (hydr)oxides. Structural characterization
droxylation at this temperature for this sample, which is even easier using 27Al NMR or X ray powder diffraction and thermal character-
when the free iron is extracted, as observed by a strong increase in the ization were therefore mandatory to gain this goal.
contribution intensity for BC-700. The many changes in these contribu-
tions seem to indicate that the free iron in the samples may interact (i) First, a higher content of secondary phases such as quartz and
with some clay minerals, such as kaolinite and muscovite, only on the feldspar was detected for the A samples, while higher amounts of
surface of the clay minerals, such as the surrounding kaolinite sheets. kaolinite and muscovite were observed for the B samples. For a
Finally, two hypotheses can be determined from these results: (i) the similar iron content, the darker color of the A samples after cal-
A25 sample contains more free iron phases, easily accessible to the cination at 350 °C means that the iron in these samples is more
chemical treatment, and has iron phases that interact with the musco- accessible.
vite minerals, and (ii) the B25 sample is more characterized by kaolinite (ii) A higher total shrinkage with temperature is observed for the B

7
J. Peyne et al. Applied Clay Science xxx (xxxx) xxx–xxx

Fig. 7. Theoretical model of the differences for (a, e) A25 and (b, ) B25 clays used in the bricks industry.

clays before and after the iron extraction. This result is accentuated Kubelka, P., Munk, F., 1931. Ein Beitrag Zur Optik Der Farbanstriche. Z. Tech. Phys. 12,
after the iron extraction. This arises from the presence of structural 593–601.
Li, H., Dong, L., Jiang, Z., Yang, X., Yang, Z., 2016. Study on utilization of red brick waste
iron in the kaolinite phases, which favors sintering. powder in the production of cement based red decorative plaster for walls. J. Clean.
(iii) 27Al NMR showed that some iron phases, in goethite form, inter- Prod. 133, 1017–1026.
acted with the clay minerals muscovite and kaolinite for the A and Lopez, F., Martirena Fernandez, R., Scrivener, J.F., 2011. The origin of the pozzolanic
activity of calcined clay minerals: a comparision between kaolinite, illite and mon-
B samples, respectively. tmorillonite. Cem. Concr. Res. 45, 113–222.
Mackenzie, K.J.D., Brown, I.W.M., Cardile, C.M., Meinhold, R.H., 1987. The thermal
References reactions of muscovite studied by high resolution solid state 29Si and 27Al NMR. J.
Mater. Sci. 22, 2645–2654.
Madejova, J., 2003. FTIR techniques in clay mineral studies. Vib. Spectrosc. 31, 1–10.
Ashbroook, S.E., McManus, J., MacKenzie, K.J.D., Wimperis, S., 2000. Multiple Quantum Madejova, J., Komadel, P., 2001. Baseline studies of the clay minerals society source
and Cross Polized 27Al NMR of mechanically treated mixtures of kaolinite and clays: infrared methods. Clay Clay Miner. 49, 410–432.
gibbsite. J. Phys. Chem. B 104, 6408–6446. Massiot, D., Dion, P., Alcover, K.F., Bergaya, F., 1995. 27Al and 29Si MAS NMR study of
Autef, A., Joussein, E., Poulesquen, A., Gasgnier, G., Pronier, S., Sobrados, I., 2013. Role kaolinite thermal decomposition by controlled rate thermal analysis. J. Am. Ceram.
of metakaolin dehydroxylation in geopolymer synthesis. Powder Technol. 250, Soc. (11), 2940–2945.
33–39. Mehra, O.P., Jackson, M.L., 1960. Iron oxide removal from soils and clays by a dithionite-
Bardy, M., Bonhomme, C., Fritsch, E., Maquet, J., Hajjar, R., Allard, T., Derenne, S., Calas, citrate system buffered with sodium bicarbonate. Clay Clay Miner. 317–327.
G., 2007. Al speciation in tropical podzols of the Upper Amazon Basin: a solid state Newman, R.H., Childs, C.W., Churchman, G.J., 1994. Aluminum coordination and
27
Al MAS and MQMAS NMR study. Geochim. Cosmichim. Ac. 71, 3211–3222. structural disorderin halloysite and kaolinite by 27Al NMR spectroscopy. Clay Miner.
Barron, V., Torrent, J., 1986. Use of the Kubelka Munk theory to study the influence of 29, 305–312.
iron oxides on soil colour. Eur. J. Soil Sci. 37, 499–510. Oikonolopoulis, I.K., Perraki, M., Tougiannidis, N., Perraki, T., Kasper, H.U., Gurk, M.,
Benharrats, N., Belbachir, M., Legrand, A.P., D'espinose de la Caillerie, J.B., 2003. 29Si 2015. Clays from Neogene Achlada lignite deposits in Florina basin (Wester
and 27Al MAS NMR study of the zeolitization of kaolin by alkali leaching. Clay Miner. Macedonia, N. Greece): a prospective resource for the ceramics industry. Appl. Clay
38, 49–61. Sci. 103, 1–9.
Bergaya, F., Lagaly, G., 2013. Chapter 1 - General Introduction: Clays, Clay Minerals, and Pardal, X., Brunet, F., Charpentier, T., Pochard, I., Nonat, A., 2012. 27 Al and 29Si Solid
Clay Science, Developments in Clay Science. vol. 5. pp. 1–19. State NMR characterization of calcium aluminosilicate hydrate. Inorg. Chem. 5,
Carlos Duarte Cavalcante, L., Sousa Bezera da Silva, H.L., Domingos Fabris, J., Domingos 1827–1836.
Ardisson, J., 2017. Red and Yellow Ocres from the Archaleogical Site Pedra to Parker, T.W., 1969. A classification of kaolinites by infrared spectroscopy. Clay Miner. 8,
Cantagalo I, in Piripiri, Piaui, Brazil. In: Hyperfine Interact. 2017. pp. 222–238. 135–147.
Carroll, D.L., Kemp, T.F., Bastow, T.J., Smith, M.E., 2005. Solid state NMR character- Peyne, J., Gautron, J., Doudeau, J., Joussein, E., Rossignol, S., 2017. Influence of calcium
ization of the thermal transformation of a Hungarian white illite. Solid State Nucl. addition on calcined brick clay based geopolymers: a thermal and FTIR spectroscopy
Mag. 28, 31–43. study. Constr. Build. Mater. 152, 794–803.
Castelein, O., Soulestin, G., Bonnet, J.P., 2000. Influence of heating rate on the thermal Ravinska, R., et al., 2010. Mineralogical characterization studies of ancient potteries of
comportment and the mullite formation from a kaolin raw material. Ceram. Int. Tamilnadu, India by FTIR spectroscopy. J. Chem-NY 7, 185–190.
27 (5). Ruan, H.D., Frost, R.L., Kloprogge, J.L., Duong, L., 2002. Infrared spectroscopy of goethite
Cornell, R.M., Schwertmann, U., 2002. The Iron Oxides, Structure, Properties, Reactions, dehydroxylation: FTIR microscopy of in situ study of the thermal transformation of
Occurrences and Uses. Wiley VCH Verlag GmbH & Co, KGa, Weinheim. goethite to hematite. Spectrochim. Acta A 58, 967–981.
Felhi, M., Tlili, A., Gaied, M.E., Montacer, M., 2008. Mineralogical study of kaolinitic Sanchez-Munoz, L., Sanz, J., Sobrados, I., Gan, Z., 2013. Medium range order in dis-
clays from Sidi El Bader in the far North of Tunisia. Appl. Clay Sci. 39, 208–217. ordered K feldspars by multinuclear NMR. Am. Mineral. 98, 2115–2131.
Gualtieri, A.F., Ricchi, A., Lassinantti, M., Gualtieri, S., Maretti, M., 2015. Kinetic study of Schroeder, P.A., Pruett, R.J., 1996. Fe ordering in kaolinite: insights from 29Si and 27Al
the drying process of clay bricks. J. Therm. Anal. Calorim. 153–167. MAS NMR spectroscopy. Am. Mineral. 87, 26–38.
Guzlena, S., Sakale, G., Certoks, S., 2017. Clayey material analysis for assessment to be Sigg, J., 1991. Les produits de terre cuite. pp. 135 (Edition Septima – Paris).
used in ceramic building materials. Modern Building Materials, Structures and Soro, N.S., 2003. Influence des ions fer sur les transformations thermiques de la kaolinite.
Techniques. MBMST 2016. Process. Eng. 172, 333–337. In: Thèse de l'Université de Limoges, Spécialité: Sciences des Matériaux Céramiques
Huang, C.K., Kerr, P.F., 1960. Infrared study of the carbonate minerals. Am. Mineral. 45, et Traitements de Surface.
311–324. Stanmore, B.R., Gilot, P., 2005. Review-calcination and carbonation of limestone during
Jimenez-Quero, V., Maza-Ignacio, O.T., Guerrero-Paz, J., Campos-Venegas, K., 2017. thermal cycling for CO2 sequestration. Fuel Process. Technol. 86, 1707–1746.
Industrial wastes as alternative raw materials to produce ecofriendly fired bricks. J. Takahashi, T., Kanehashi, K., Saito, K., 2008. First evidence of multiple octahedral Al sites
Phys. 792 (012065). in Na-montmorillonite by 27Al multiple quantum MAS NMR. Clay Clay Miner. 56 (5),
Kim, J., Illot, A.J., Middlemiss, D.S., Chernova, N.A., Pinney, N., Morgan, D., Grey, C.P., 520–525.
2015. 2H and 27Al solid state NMR study of the local environments in Al doped 2-line Thompson, J.G., 1984. 29Si and 27Al nuclear magnetic resonance spectroscopy of 2:1 clays
ferrihydrite, goethite and lépidocrocite. Chem. Mater. 27, 3966–3978. minerals. Clay Miner. 19, 229–236.
Kornmann, M., 2009. Matériaux de terre cuite – Propriétés et produits, c906. Techniques Vaculikova, L., Plevova, E., Vallova, S., Koutnik, I., 2011. Characterization and
de l'Ingénieur.

8
J. Peyne et al. Applied Clay Science xxx (xxxx) xxx–xxx

differentiation of kaolinites from selected Czech deposits using infrared spectroscopy mechanisms and stability of binary systems containing goethite and kaolinite. Soil
and differential thermal analysis. Acta Geodyn. Geomater. 18, 59–67. Sci. Soc. Am. J. 76, 389–398.
Vodyanitskii, Y.N., 2010. Iron hydroxides in soils: a review of publications. Eurasian Soil Wel-Lin Shao, W., 1989. NMR Studies of Amorphous Silicon and Aluminosilicate Glass.
Sci. 43, 1244–1254. Iowa State University (PhD thesis).
Wanabe, T., Shimiz, H., Nagasawa, K., Masuda, A., Saito, H., 1987. 29Si and 27Al MAS Woessner, D.E., 1989. Characterization of clay minerals by 27Al nuclear magnetic re-
NMR study of the thermal transformation of kaolinite. Clay Miner. 22, 37–48. sonance spectroscopy. Am. Mineral. 74, 203–215.
Wei, S., Tan, W., Zhao, W., Zu, Y., Liu, F., Koopal, K., 2012. Microstructure, interaction

You might also like