You are on page 1of 7

J. Am. Ceram. Soc.

, 96 [11] 3589–3595 (2013)


DOI: 10.1111/jace.12535
© 2013 The American Ceramic Society

Journal
Raman Spectroscopy of Anhydrous and Hydrated Calcium
Aluminates and Sulfoaluminates
David Torrens-Martın,‡,† Lucia Fernandez-Carrasco,‡ Sagrario Martınez-Ramırez,§ Jordi Ib
an~ez,¶
Lluis Artus, and Thomas Matscheik


Department of Architectural Technology I, Universitat Politecnica de Catalunya, Barcelona, Spain
§
Instituto de Estructura de la Materia, IEM-CSIC, Madrid, Spain

Institut de Ciencies de la Terra Jaume Almera (CSIC), Barcelona, Spain
k
Holcim Technology Ltd, Innovation - Research and Development, Holderbank, Switzerland

Recent investigations have revealed the great potential of gypsum, bauxite, and limestone as raw materials. By burning
Raman spectroscopy for the characterization of clinker miner- of the raw materials at ~1300°C, a clinker based on the main
als and commercial Portland cements. The usefulness of this minerals C4A3S (yeelimite), C2S, and partially CS (anhydrite)
technique for the identification of anhydrous, hydrated, and is formed. Depending on the used raw materials, phases such
carbonated phases in cement-based materials has been demon- as C4AF (brownmillerite), C3A, C12A7 (mayenite), and C2AS
strated. In the present work, the application of micro-Raman can be found in minor amounts.4
spectroscopy for the characterization of the main clinker C3AH6 (hydrogarnet, also known as katoite) and AH3
phases of calcium aluminate cements and calcium sulfoalumi- (gibbsite) are the main stable hydration products of CAC.5
nate cement is explored. The main stable hydrated phases as In initial stages of the hydration reaction, the developed
well as several important carbonated phases are investigated. hexagonal phases CAH10 and/or C2AH8 may decompose
Raman measurements on the following phases are reported: onto C3AH6 and gibbsite depending on the service tempera-
(i) pure, unhydrated phases: CA, C12A7, CA2, C2AS, cubic- ture of the concrete.
C3A, C4AF, and C4A3S;  (ii) hydrated phases: ettringite, Ettringite is the main crystalline phase formed during the
monosulfoaluminate, and hydrogarnet (C3AH6); (iii) carboa- hydration of CSA. Ettringite exhibits hexagonal structure built
luminate phases: hemicarboaluminate and monocarboalumi- up of columns of [Ca6Al2(OH)1224H2O]6+ oriented along the
nate. The present results, which are discussed in terms of the z-axis; channels rise in the intercolumn space, filled with sul-
internal vibrational modes of the aluminate, carbonate, and fate groups and crystallization water.6
sulfate molecular groups as well as stretching O–H vibrations, The exposure to CO2 and/or the addition of limestone to
show the ability of Raman spectroscopy to identify the main the cements (CAC, CSA as well as OPC) leads to the forma-
hydrated and unhydrated phases in the aluminate and sulfoalu- tion of carboaluminate type AFm phases (e.g., hemi- and/or
minate cements. The Raman spectra obtained in this work pro- monocarboaluminate). In general, it may prove very difficult
vide an extended database to the existing data published in the to detect these phases by means of traditional analytical
literature. methods, such as X-Ray Diffraction (XRD) or Infrared (IR)
absorption spectroscopy. Thus, the availability of additional
I. Introduction characterization tools to unambiguously identify these phases
would be highly desirable.
C ALCIUM aluminate cements (CACs) were originally devel-
oped as cement with improved sulfate resistance
compared with portland cement. Both CAC and portland
Several studies dealing with characterization of portland
and aluminate cements have been published so far. Dyer
et al.7 explored by means of FT-Raman (laser excitation of
cements are based on calcium, silicon, aluminum, and iron
1064 nm) the C3S, C2S, and C3A pure phases and attributed
oxide.1 However, the composition of the two cements within
the low signal-to-noise ratio of their spectra to high optical
the CaO–SiO2–Al2O3 ternary system is quite distinct. CAC is
fluorescence. Newman et al.8 used as excitation source a Nd:
often used for the production of refractories. CACs with
YAG laser (1064 nm) to study cement-based systems and
low-grade alumina (40%–60% Al2O3) are generally produced
concluded that the intensity of fluorescence decreases with
by melting bauxite and limestone. Their main reactive phase,
increasing degree of hydration; moreover, suggested the near-
responsible for the strength properties of the material, is
infrared excitation as the adequate to study hydrated cement
anhydrous calcium aluminate (CA). Other minor mineralogi-
phases.
cal phases found in CACs are C2S (belite), C2AS (gehlenite),
Bonen et al.9 not only studied synthetic phases but also
and the ferrite solid solution.2 Refractory cements may
commercial portland cement clinkers. In that work, it was
contain also C12A7, CA, and CA2.3
shown that the presence of impurities can modify the posi-
Calcium sulfoaluminate cement (CSA) is a binding mate-
tion of the Raman bands arising from the silicate phases.
rial used for rapid setting, early high strength development,
Such impurities may distort the crystalline structure of the
and shrinkage compensation. CSA requires essentially
silicates and account for the observed differences between the
Raman spectra of synthetic and commercial clinkers. Raman
spectroscopy was also employed to study the structure of
H. Jennings—contributing editor
C–S–H gel. Kirkpatrick et al.10 showed the similarity of the
structure of certain C–S–H to that of tobermorite. Garbev
et al.11 used a 532-nm laser excitation to investigate the
Manuscript No. 32739. Received February 19, 2013; approved July 2, 2013. structural variations in C–S–H gel as a function of the Ca/Si

Author to whom correspondence should be addressed. e-mail: david.torrens@upc.edu ratio. These authors found that changes in the Ca/Si ratio
3589
3590 Journal of the American Ceramic Society—Torr
ens-Martın et al. Vol. 96, No. 11

lead to systematic changes in the spectra, showing that rele- (2) Analytical Methods
vant structural information can be gained from the Raman X-Ray Diffraction was performed on flat samples by using a
experiments. Philips PW-1700 powder diffractometer (Ettlingen, Germany)
Raman scattering has also been used to probe the hydra- (Bragg–Brentano geometry) equipped with a sealed Cu X-ray
tion of portland cement phases. Iban ~ez et al.12 studied the source, a secondary graphite monochromator to filter out the
hydration and carbonation of C3S and C2S and confirmed CuKb radiation, and a scintillation detector. The measure-
the fast and slow hydration reactions of C3S and C2S, respec- ments were acquired with standard operating conditions of
tively. Black et al.13 investigated the hydration of C3A and 40 kV and 50 mA. The angular range of the measurements
C4AF in the presence and absence of sulfate compounds. was 5°–60° (in 2h). The XRD patterns of the powders of
These authors observed that the hydration of C3A is sizably CA, C12A7, CA2, cubic-C3A, C4AF, C2AS, and C4A3S
faster than that of C4AF. In both cases, however, the early showed that the samples did not contain any secondary com-
hydration stages gave rise to C4AH19. With regard to the pound. In contrast, we found that some of the AFm sample
carbonation of these phases, the evolution of C3A to hemi- contained small proportion of hydrogarnet and some port-
and monocarboaluminate, with formation of aragonite, was landite in the case of C4AH13.
observed. In contrast, C4AF only gave rise to a hemicarbon- Micro-Raman measurements were performed at room tem-
ate phase together with calcite. The hydration in presence of perature with two different systems: (i) a Renishaw (Marne la
sulfate was observed to produce ettringite and different forms Valle, France) inVia micro-Raman system equipped with a Le-
of AFm. Frias et al.14 used Raman spectroscopy to investi- ica microscope, a charge-coupled device (CCD) camera, and a
gate the hydration of white cement pastes blended with 532-nm laser with an output power of 10 mW. Typical spectra
metakaolin. from 90 to 4000 cm1, with a resolution of 4 cm1, were
Synthetic ettringite and monosulfoaluminate were studied recorded with this system. Five 10-s scans were averaged to
by Renaudin et al.15 The authors focused on the vibrational improve the signal-to-noise ratio. (ii) a Jobin-Yvon T64000 tri-
modes of the sulfate group and the OH bonds. They ple spectrometer (Horiba, Barcelona, Spain) equipped with a
observed a shift in the symmetric stretching (m1) of the microscope, a CCD camera, and a 514.5-nm Ar laser. Typical
sulfate group, from 990 cm1 in ettringite up to 982 cm1 in excitation powers were in the range 50–100 mW. The resolu-
the monosulfoaluminate. A shift in the stretching vibrations tion of the measurements, acquired with the double subtrac-
of the OH bond were also observed, from 3638 cm1 in et- tive configuration, was around 2 cm1. Long integration times
tringite up to 3688 cm1 in the monosulfoaluminate. The were considered (30 min).
structural changes in the ettringite phase by partial substitu- The correct calibration of both instruments was verified by
tion of Al3+ for Fe2+ was also investigated. More recently, comparing the frequency position of the Stokes and
Mesbah et al.16 were able to identify Cl and CO32 substi- anti-Stokes bands and/or by checking the position of the Si
tutions in AFm phases. In turn, Martinez-Ramirez et al.17 first-order peak at 520.6 cm1. To check for the consis-
identified an AFt phase formed in the hydration of CSA. tency of the results, Raman spectra were collected on differ-
This compound decomposes after carbonation and yields ent sample portions, which were found to yield similar
rapidcreekite, a hydrated calcium sulfate–carbonate com- results in all samples investigated.
pound.
McMillan et al.18 presented a Raman study about CA
glasses and crystals. They concluded that the Raman spec- III. Results and Discussion
trum of CaAl2O4 glasses may be interpreted in terms of a The Raman measurements on crystalline and amorphous
fully polymerized network of tetrahedral aluminate units, aluminates published in several of these studies9,11–19 have
which is depolymerised on addition of CaO as occurs in been employed to investigate the samples of the present work.
binary silicate systems. Gastaldi et al.19 used Raman spec-
troscopy, in combination with XRD measurements, to inves-
tigate the hydration of CSA: the mineral phases responsible (1) Raman Spectra of Anhydrous Powders
for the observed Raman bands were identified and the hydra- Figures 1–4 show the Raman spectra of the anhydrous
tion process was monitored. aluminate phases studied in this work. Table I displays the
In this work, Raman spectroscopy was used to study sev- wave numbers position of the main Raman peaks observed
eral anhydrous, hydrated, and carbonated calcium aluminate in the different samples together with the band assignments.
and sulfoaluminate phases. We discuss the results in terms of Figure 1 shows the Raman spectrum of CA, dominated by
the frequencies of the different vibrational modes of the a strong, sharp signal at 522 cm1 and a weaker band at
silicate, sulfate, and carbonate molecular groups and provide
an updated database of Raman-scattering data for aluminate
and sulfoaluminate cements. Particular emphasis is placed
with regard to the high-frequency regions, up to 4000 cm1,
corresponding to the stretching modes of OH bonds, which
are relevant for the study and identification of the hydrated
phases. This spectral range has been disregarded in many
previous works.

II. Materials and Methods


(1) Materials
Stoichiometric proportions of aluminum oxide, calcium oxide
and, when appropriate, anhydrous calcium sulfate were used
in the synthesis of CA, C12A7, CA2, cubic-C3A, C4AF,
C2AS, and C4A3S.  The powders were hand-grinded, pres-
sure-pelletized, and heated in a platinum crucible at tempera-
tures ranging from 1200°C to 1400°C.
Details about the preparation of the hydrated phases,
ettringite, monosulfoaluminate, hydrogarnet, hydroxy-AFm,
hemicarboaluminate, and monocarboaluminate can be found
in Ref. [20] Fig. 1. Raman spectrum of CA.
November 2013 Raman Spectroscopy Calcium Aluminates 3591

Table I. Calcium Aluminates Raman Absorption Bands


cm1

CA CA2 C3A C12A7

AlO45 m1 522 s/546 w 412 s/458 508 521


m3 793 661/689 s 757 772
vw, very weak; s, strong; m, medium; b, broad; w, weak; sh, shoulder; vs,
very strong.

AlO45 and Ca2+ ions show up in the spectrum. In particu-


lar, a more prominent peak is visible in the region close to
144 cm1.
In contrast, the Raman spectrum of CA2 (Fig. 2) exhibits
many different features that can be mainly ascribed to two
different groups of bands. The first group, located between
640 and 950 cm1 with two bands of comparable intensity at
661 and 689 cm1, can be attributed to m3 modes of the
Fig. 2. Raman spectrum of CA2. AlO45 tetrahedra. The second group of bands lies
below 500 cm1 with a strong, sharp peak at 412 cm1 and
a weaker feature at 458 cm1. These two bands probably
arise from m1 modes of the AlO45 tetrahedra and, together
with the peaks that are observed in the high–wave number
spectral region (up to 950 cm1), provide a fingerprint for
the presence of CA2 in the aluminate cements.
The Raman spectrum of our C12A7 sample is depicted in
Fig. 3. In this case, the spectrum is dominated by two Raman
bands, one at 521 cm1 and another one at 772 cm1. These
two features can be assigned to m1 and m3 modes of the
AlO45 groups, respectively. The absence of an additional
feature in the high-frequency shoulder of the m1 peak,
together with the lower frequency of the m3 peak and the pres-
ence of a relatively intense band at around 300 cm1 allows
one to discriminate between the presence of this phase and
that of CA. The band at 300 cm1 and the Raman features
that appear at lower wave numbers can be assigned to Ca–O
vibrations.
Figure 4 shows the Raman spectrum of cubic C3A, which
is dominated by two sharp bands at 508 and 757 cm1. The
C3A structure is based on isolated six-membered rings of
Fig. 3. Raman spectrum of C12A7. AlO45 tetrahedra.13 These rings are formed by two corner-
sharing oxygens per tetrahedron, giving rise to a structure
with two nonbridging oxygen per AlO45 group. These two
tetrahedra can be considered as AlO2 units. The strong
Raman band at 757 cm1 may be associated with a symmet-
ric stretching vibration of these AlO2 groups. The weaker
Raman bands and the strong absorption in this region may
be related to asymmetric aluminate stretching motions.
The Raman peak at 508 cm1 may be associated with m1
AlO45.15 These signals are characteristic of cubic C3A.13–16

Fig. 4. Raman spectrum of C3A.

546 cm1. This doublet, which can be assigned to symmetric


stretching (m1) of the Al–O bonds in the AlO45 tetrahedra,
allows one to unambiguously distinguish the presence of this
phase in aluminate cements. The broad band at 793 cm1
can be attributed to asymmetric stretching modes (m3) of the
AlO45 groups. In the low–wave number region of the spec-
trum, below 300 cm1, several external lattice modes of Fig. 5. Raman spectrum of C4AF.
3592 Journal of the American Ceramic Society—Torr
ens-Martın et al. Vol. 96, No. 11

The sample also presents a weak band at 360 cm1 which frequency position of the 531 cm1 peak is consistent with m1
probably arises from Ca–O vibrations. modes from AlO45. In the low-frequency region, a sharp
In contrast, the Raman spectrum of C4AF reveals the band at 307 cm1 arising from internal modes of the alumi-
poor crystallinity of the sample used in this study (Fig. 5). nate or sulfate ionic groups is particularly prominent.
Two broad bands show up in this spectrum: the first one In the case of C4A3S (Fig. 7), a very strong and sharp
situated around 280 cm1, which can be ascribed to symmet- band at 993 cm1 arising from m1 SO42 vibrations domi-
ric (m2) and/or antisymmetric (m4) bending modes of the nates the Raman spectrum. Additional weaker bands at 1201
O–Al–O (or O–Fe–O) bonds in the [(Fe,Al)O4]5 or [(Fe,Al) and 614 cm1, probably arising from m3 and m4 modes of
O6]9 groups; the second band is located around 770 cm1 SO42,19 also appear in the spectrum. The spectrum reveals
and can be assigned to [(Fe,Al)O4]5 or [(Fe,Al)O6]9 that the wave number position of vibrational modes from the
stretching vibrations.7,16,18 AlO45 groups is lower in this sample: m3 and m1 modes are
Figure 6 shows the Raman spectrum of C2AS. Several found at 645 and 482 cm1, respectively. However, these
groups of bands can be identified in this spectrum. Four bands are fairly weak and broad, and as a consequence it is
features are visible in the high-frequency spectral region of fairly difficult to use these features for identification pur-
the spectrum: a weak and relatively narrow band at poses. The fact that the Raman spectrum of C4A3S only
799 cm1 and a weak and broader feature at 853 cm1, both exhibits a single sharp peak at 993 cm1 suggests that alter-
of which can be attributed to m1 modes of the SiO44 groups. native analytical techniques should be used to unambigu-
Also, a peak of medium intensity at 916 cm1 and a sharper ously identify the presence of C4A3S in the aluminate
band at around 980 cm1 are visible, both of which are cements.
assigned to m3 modes of SiO44. On the other hand, two
Raman bands emerge at lower frequencies: a strong and
sharp peak at 631 cm1 and a weaker band around (2) Raman Spectra of Hydrated Powders
658 cm1 arising from m3 vibrations from AlO45 tetrahedra. The Raman spectra corresponding to synthetic hydrated CA
This doublet above 630 cm1 provides a fingerprint for the samples are displayed in Figs. 8–11 and Tables I–V.
presence of C2AS and allows one to unambiguously identify Raman spectra of ettringite and monosulfoaluminate are
this phase from the Raman spectra. Additional weak bands shown in Fig. 8. Both spectra exhibit a strong Raman signal
emerge in the spectrum at 411, 462, and 480 cm1. These around 980 cm1 due to m1 SO42 vibrations as well as a
features can be assigned to m2 SiO44 modes, while the weak feature at 616 cm1 due to m4 SO42 vibrations.8 Both

Fig. 6. Raman spectrum of C2AS. Fig. 8. Raman spectra of ettringite and monosulfoalumante.

Fig. 7. 
Raman spectrum of C4A3S. Fig. 9. Raman spectrum of hydroxy-AFm.
November 2013 Raman Spectroscopy Calcium Aluminates 3593

Table III. Ettringite and Monosulfoaluminate Raman


Absorption Bands
cm1

Ettringite Monosulfoaluminate

C3A3cSH32 C4ASH 12

Al(OH)6 534 525


SO42
m1 980 s 980 s
m3
m4 616 w 616 w
stretching O–H
3458 b
3557 sh
3619 b
3641 b
Fig. 10. Raman spectrum of C3AH6. 3692 s
vw, very weak; s, strong; m, medium; b, broad; w, weak; sh, shoulder; vs,
very strong.

Table IV. Katoite and Hydroxy-AFm Raman Absorption


Bands
cm1

Hydrogarnet AFm
C3AH6 C4AH13–19 Portlandite

Al(OH)6
m1 327 334w/357w
m3 521
stretching O–H
3620 3620
3651 s 3652 b
3673 vw
3694 vw
vw, very weak; s, strong; m, medium; b, broad; w, weak; sh, shoulder; vs,
Fig. 11. Raman spectra of hemicarboaluminate and monocarboa- very strong.
luminate.

Table V. Hemicarboaluminate and Monocarboaluminate


Table II. C4AF, C2AS, and C4A3S Raman Absorption Bands Raman Absorption Bands

cm1 cm1

C4AF C4A3S C2AS Hemicarboaluminate Monocarboaluminate


C4A
cH11 C4Ac0.5H12
SiO44
m1 799/853 Al(OH)6
m2 411/462/480 531 s 531
m3 916/980 CO32
AlO45 1068 s 1068 vw
m1 482 531 1086 w 1086
m3 645 631/658 stretching O–H
SO42 3542 b
m1 993 s 3592 sh
m3 1201 3615 sh
m4 614 3627
[(Fe,Al)O45] 280 b 3659 sh
[(Fe,Al)O69] 770 b 3670 s
3679 s
vw, very weak; s, strong; m, medium; b, broad; w, weak; sh, shoulder; vs, 3683 b
very strong.
vw, very weak; s, strong; m, medium; b, broad; w, weak; sh, shoulder; vs,
very strong.
spectra can be distinguished due to the different wave vector
position of the characteristic band arising from Al(OH)6
stretching vibrations. In the case of monosulfoaluminate, this and 3800 cm1, two broad bands emerge in the Raman spec-
band is fairly sharp and appears at 534 cm1, whereas in the trum of ettringite: a first, broad band around 3458 cm1,
case of ettringite it is much weaker and shows up at sizably and a second narrower band around 3641 cm1. In the case
higher wave numbers. In the high–wave number region of the monosulfoaluminate sample, a broad band about
corresponding to O–H stretching vibrations, between 3200 3619 cm1 with a shoulder centered at 3557 cm1 and a
3594 Journal of the American Ceramic Society—Torr
ens-Martın et al. Vol. 96, No. 11

narrow, intense band at 3692 cm1 are visible in the Raman vibrational interactions with neighboring groups, in relation
spectrum. These Raman spectra reveal that the features to the characteristic vibrational modes of AlO4 tetrahedra
appearing in the high–wavenumber region of O–H stretching and AlO6 octahedra in a series of aluminate compounds. The
vibrations may be highly informative to identify the presence vibrational modes associated with the internal modes of the
of hydrated aluminates in CAC and CSA. [AlO45] complex ion have been identified in the Raman
In the case of the hydroxy-AFm (Fig. 9), several bands at spectra of different CAs.22–27
3620 (sharp), 3652 (broad), 3673 (very weak), and 3694 cm1 Spectra of CA, C12A7, and C3A show a similar profile, but
(very weak) are observed. It should be noted that, as there is a difference in the relative intensity of m1 and m3 of
revealed by XRD, this sample contains sizable amounts of AlO45 absorption bands. When the Ca/Al ratio is low, the
hydrogarnet and portlandite [see also Ref. (20)]. Given that m1/m3 relative intensity was found higher; as this ratio grows,
the O–H stretching bands of hydrogarnet and portlandite, the relative intensity decreases: in the C3A spectrum, m3 is
respectively, emerge at 3651 and 3620 cm1, the observed higher than m1. However, in the case of CA2 spectrum, there
peaks at these wave number positions may be also assigned are many bands for m1 and also m3. According to Tarte,21 this
to these two phases. However, in the case of the intense peak fact is due to the Al neighbors producing an environment
at 3620 cm1 that dominates the Raman spectrum of around each AlO4 rich in Al that provoke several signals.
hydroxy-AFm, this peak can be unambiguously assigned to In the hydrated aluminate phase spectra of ettringite and
the majority AFm phase. In the low–wave number spectral monosulfoaluminate, we have two different structures. The
region, bands arising from hydrogarnet at 327 (medium) and structure of ettringite is formed by columns of calcium oxide
521 cm1 (strong) are also observed (Fig. 10). These bands and aluminum with hydroxyl anions where the sulfate is
are originated by stretching vibrations of the Al(OH)6 located in the channels between these columns. Monosulfoa-
groups.7,17 luminate structure is laminar, with layers formed for Ca and
At lower frequencies, hydroxy-AFm exhibit a somewhat Al and layers interspersed of sulfate. Raman spectra are simi-
lower intensity as expected from the lower relative amount of lar in the sulfate bands, but they present the main differences
hydrogarnet, weak bands at 210, 250, 390, and 405 cm1 as in the OH zone, indicating different structures, laminar, and
well as a somewhat sharper band at 358 cm1. The 250 and columns.
358 cm1 bands can be attributed to portlandite, the rest of Respecting the carbonated phases, the mCA and hCA
weak, broad features provide additional signatures to identify have the same bands for the Al and the carbonate groups.
the presence of hydroxy-AFm. The difference in the spectra is due to the band intensity,
Raman measurements on hCA and mCA samples yield being lower for the hCA, due to the presence of hydroxyl
the spectra of Fig. 11. The main features that appear in the anions. The two phases have a laminar structure, but the
low–wave number region for these two compounds arise carbonate presence is higher in the mCA.
from stretching vibrations from octahedral Al(OH)6 and
stretching vibrations from the carbonate groups (CO32).
IV. Conclusions
The band corresponding to Al(OH)6, centered around
531 cm1, is stronger in the case of mCA. The features We have carried out Raman-scattering measurements on
arising from the carbonate groups appear at similar wave several anhydrous and hydrated CA cement phases. In the
numbers for both compounds, but with different relative case of the hydrated phases, particular attention has been
intensities. The mCA sample exhibits a strong peak around paid to the high–wave number spectral range corresponding
1068 cm1 and a weaker band at 1086 cm1.In contrast, the to stretching O–H vibrations. The Raman peaks observed for
spectrum from hCA exhibits a much weaker band at the different phases have been assigned in terms of the inter-
1068 cm1, whereas the peak centered at 1086 cm1 domi- nal vibrational modes of the aluminate, carbonate, and
nates the spectrum. The peak at 1068 cm1 is attributed to a sulfate molecular groups as well as stretching O–H vibra-
carbonate group directly bonded to the main layer (i.e., tions. Our data confirm the high sensitivity of Raman spec-
linked to Ca2+ cations), whereas the latter is believed to arise troscopy to identify the main hydrated and unhydrated
from weakly bonded carbonate groups situated at the center phases in the aluminate and sulfoaluminate cements. In the
of the interlayer.9 The relative intensities of these two peaks particular case of mineral phases that are difficult to distin-
allow one to unambiguously discriminate the presence of guish by means of XRD measurements, Raman spectroscopy
these two compounds in the aluminate cements. With regard emerges as a highly effective tool for the identification
to the high–wave number peaks, the Raman spectrum of of these phases. The Raman data reported in the present
mCA indicates that the degree of crystallinity of this phase is work allow one to extend the existing Raman database
much higher than that of hCA. Several bands at 3542 published in the literature dealing with cements and related
(broad), 3615 (shoulder), 3627, 3670 (strong), and 3679 cm1 compounds.
(strong) emerge in the spectrum of mCA. In contrast, the
spectrum of hCA exhibits a broad band centered at
Acknowledgments
3683 cm1, with shoulders around 3659 and 3592 cm1.
The Raman bands tend to shift to higher wave numbers in This study was funded by the MINECO (Ministerio de Economıa y Competi-
samples containing higher amounts of calcium oxide in the tividad) through BIA0767-2008 project.
formula unit (Tables I to V). In contrast, when the amount
of aluminum oxide in the formula unit increases, the wave References
number of the bands tends to decrease. With regard to the
1
Raman bands associated with the sulfate groups, it is found K. L. Scrivener, “Calcium Aluminate Cement”;pp. 64–93 in Advanced
Concrete Technology, Constituent Materials, Edited by J. Newman and B. S. Choo.
that these peaks display sizable intensity variations with Butterworth-Heineman, Oxford, 2003.
composition, whereas their wave number position remains 2
B. Touzo, A. Glotter, and K. L. Scrivener, “Mineralogical Composi-
more or less unchanged. The combination of (i) the wave tion of Fondu Revisited”; pp. 129–38 in Calcium Aluminate Cements 2001,
number position of the aluminate bands, (ii) the intensity of R. J. Mangabhai and F. P. Glasser. IOM Communications, London, 2001.
3
J. Bensted, Calcium Aluminate Cements, Structure and Performance of
the sulfate bands, and (iii) the shape, number of bands, and Cements, 2nd edition, Edited by J. Bensted and P. Barnes. Taylor and Francis,
wave number position of the O–H stretching modes may be London, 2002.
4
highly informative to identify the presence of anhydrous and Z. Peixing, C. Yimin, S. Liping, Z. Guanying, H. Wenmei, and W. Jianguo,
hydrated aluminate phases in cement materials. 9th International Congress on the Chemistry of Cement. p. 201, National Coun-
cil for Cement and Building Materials, New Delhi, 1992.
Tarte21 discussed on the factors influencing the vibrational 5
K. L. Scrivener and A. Capmas, “Calcium Aluminate Cements”; pp.
frequencies of cation-oxygen coordinated groups, namely the 713–82 in Chemistry of Cement and Concrete, Edited by C. Hewlett. Butter-
value of the coordination number, ordinated groups, and worth-Heinemann, Oxford, 2006
November 2013 Raman Spectroscopy Calcium Aluminates 3595
6 17
A. E. Moore and H. F. W. Taylor, “Crystal Structure of Ettringite,” Acta S. Martınez-Ramırez and L. Fernandez-Carrasco, “Carbonation of
Cryst. B, 26, 386–93 (1970). Ternary Cement System,” Constr. Buil. Mater., 27 [1] 313–8 (2012).
7 18
C. D. Dyer, P. J. Hendra, and W. Forsling, “The Raman Spectroscopy of P. McMillan and B. Piriou, “Raman Spectroscopy of Calcium Aluminate
Cement Minerals Under 1064 nm Excitation,” Spectrochim. Acta, 49A [5/6] Glasses and Crystals,” J. Non-Cryst. Solids, 55 [2] 221–42 (1983).
19
715–22 (1993). D. Gastaldi, E. Boccaleri, F. Canonico, and M. Bianchi, “The use of
8
S. P. Newman, S. J. Clifford, P. V. Coveney, V. Gupta, J. D. Blanchard, Raman Spectroscopy as a Versatile Characterization Tool for Calcium
F. Serafin, D. Ben-Amotz, and S. Diamond, “Anomalous Fluorescence in Sulfoaluminate Cements: A Compositional and Hydration Study,” J. Mater.
Near-Infrared Raman Spectroscopy of Cementitious Materials,” Cem. Concr. Sci., 42 [20] 8426–32 (2007).
20
Res., 35 [8] 1620–8 (2005). T. Matschei, B. Lothenbach and F. P. Glasser, “The AFm Phase in Port-
9
D. Bonen, T. J. Johnson, and S. L. Sarkar, “Characterization of Principal land Cement,” Cem. Concr. Res., 37 [2] 118–30 (2007).
21
Clinker Minerals by FT-Raman Microspectroscopy,” Cem. Concr. Res., 24 [5] P. Tarte, “Infrared Spectra of Inorganic Aluminates and Characteristic
959–65 (1994). Vibrational Frequencies of AlO4 Tetrahedra and AlO6 Octahedra,” Spectrochi-
10
R. James Kirkpatrick, J. L. Yarger, P. F. McMillan, P. Yu, and X. Cong, mica Acta: Part A Mol. Spectrosc., 23A [7] 2127–43 (1967).
22
“Raman Spectroscopy of C-S-H,Tobermorite, and Jennite,” Advn. Cem. Bas. J. Bensted, “An Infrared Spectral Examination of Calcium Aluminate
Mat., 5 [9] 3–99 (1997). Hydrates and Calcium Aluminate Sulphate Hydrates Encountered in Portland
11
K. Garbev, P. Stemmermann, L. Black, C. Breen, J. Yarwood, and B. Cement Hydration”; Alluminati di Calcio, Seminario Internazionale Torino,
Gasharova, “Structural Features of C–S–H(I) and Its Carbonation in Air-A Italy, 1982.
23
Raman Spectroscopic Study. Part I: Fresh Phases,” J. Am. Ceram. Soc., 90 [3] C. S. Deng, C. Breen, J. Yarwood, S. Habesch, J. Phipps, B. Crasterb,
900–7 (2007). and G. Maitland, “Ageing of Oilfield Cement at High Humidity: A Combined
12
J. Ib
an~ez, Ll. Art
us, R. Cusc
o, A. L
opez, E. Menendez, and M. C. Andrade, FEG-ESEM and Raman Microscopic Investigation,” J. Mater. Chem., 12 [10]
“Hydration and Carbonation of Monoclinic C2S and C3S Studied by Raman 3105–12 (2002).
24
Spectroscopy,” J. Raman Spectrosc., 38 [1] 61–7 (2007). L. Black, C. Breen, J. Yarwood, C. S. Deng, J. Phipps, and G. Maitland,
13
L. Black, C. Breen, J. Yarwood, J. Phipps, and G. Maitland, “In Situ “Hydration of Tricalcium Aluminate (C3A) in the Presence and Absence of
Raman Analysis of Hydrating C3A and C4AF Pastes in Presence and Absence Gypsum—Studied by Raman Spectroscopy and X-ray Diffraction,” J. Mater.
of Sulphate,” Adv. Appl. Ceram., 105 [4] 209–16 (2006). Chem., 16 [13] 1263–72 (2006).
14 25
M. Frias and S. Martınez-Ramırez, “Use of Micro-Raman Spectroscopy M. Conjeaud and H. Boyer, “Some Possibilities of Raman Microprobe in
to Study Reaction Kinetics in Blended White Cement Pastes Containing Cement Chemistry,” Cem. Concr. Res., 10 [1] 61–70 (1980).
26
Metakaolin,” J. Raman Spectrosc., 40 [12] 2063–8 (2009). G. Ventkatamaran and V. C. Sahni, “External Vibrations in Complex
15
G. Renaudin, R. Segni, D. Mentel, J. M. Nedelec, F. Leroux, and C. Crystals,” Rev. Mod. Phys., 42 [4] 409–70 (1970).
27
Taviot-Gueho, “A Raman Study of the Sulfated Cement Hydrates: Ettringite L. Fernandez-Carrasco, D. Torrens-Martın, L. M. Morales, and S. Marti-
and Monosulfoaluminate,” J. Adv. Concr. Technol., 5 [3] 299–312 (2007). nez-Ramırez, “Infrared Spectroscopy in the Analysis of Building and Con-
16
A. Mesbah, C. Cau-dit-Coumes, F. Frizon, F. Leroux, J. Ravaux, and G. struction Materials”; pp. 369–82 in Infrared Spectroscopy - Materials Science,
Renaudin, “A New Investigation of the Cl- CO32 Substitution in AFm Engineering and Technology, Edited by Theophile Theophanides InTech,
Phases,” J. Am. Ceram. Soc., 94 [6] 1901–10 (2011). Rijeka, 2012. h

You might also like