Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Cement 12 (2023) 100064

Contents lists available at ScienceDirect

Cement
journal homepage: www.sciencedirect.com/journal/cement

Kilogram scale synthesis of C3A polymorphs and their hydration reactions


Daniel Axthammer a, Tobias Lange a, Joachim Dengler b, Torben Gädt *, a
a
Chair for the Chemistry of Construction Materials, TUM School of Natural Sciences, Technical University of Munich, Lichtenbergstraße 4, 85748 Garching, Germany
b
BASF Construction Additives GmbH, Dr.-Albert-Frank-Straße 32, 83308 Trostberg, Germany

A R T I C L E I N F O A B S T R A C T

Keywords: Studies on the properties of pure C3A phases are often limited to methods requiring small sample amounts due to
Ca3Al2O6 the lack of a convenient laboratory synthesis yielding sample amounts exceeding 100 g. Here, we report a simple
Synthesis and large scale lab method for the synthesis of C3A polymorphs with yields of up to 500 g per batch. Commercial
Cement
calcium aluminate cement (CAC) was used to prepare cylindrical green bodies of CaCO3 and Al2O3 (and NaNO3
Hydration
Calorimetry
for orthorhombic and monoclinic polymorphs). The green bodies were sintered at 1300 ◦ C and 1400 ◦ C
respectively. The chemical and mineralogical compositions of the obtained C3A polymorphs were analyzed by X-
ray powder diffraction and X-ray fluorescence spectroscopy. The reactivities of these C3A polymorphs were
compared to conventionally synthesized C3A (using mechanical powder compaction prior to sintering) via in-situ
isothermal heat flow calorimetry. Additionally, we demonstrate that synthetic C3A retains its reactivity over one
year if stored appropriately. As the new synthesis protocol yields hundreds of grams of C3A, it enables experi­
mental methods such as slump flow testing with pure phases, which is also reported for all polymorphs.

1. Introduction flow maximum of the silicate reaction. This is necessary to prevent a


disturbance of the aluminate to silicate balance in the pore solution,
The very early stage of ordinary Portland cement (OPC) hydration which is known to lead to strong retardation of the silicate reaction [10,
has a strong influence on the engineering properties of cementitious 11,16–18]. Hence, it is highly relevant to study the reactivity of different
materials such as concrete. The rheological behavior [1–4], the setting C3A polymorphs both in the presence and absence of sulfates (or other
time [5,6] or early age shrinkage [7–9] are linked to this early reactivity. additives). Unfortunately, the multiphase character of Portland cement
Importantly, the phase with the fastest initial hydration reaction in renders reactivity studies on individual phases such as C3A very difficult.
Portland cement is tricalcium aluminate C3A. It typically accounts for While often polyphasic model clinkers with a reduced phase content
4-10 wt.% of the mass and is present either as cubic or orthorhombic C3A compared to commercial Portland cement are a preferred option for
polymorph or a mixture of both [10,11]. The incorporation of increasing model hydration studies due to the specific distribution of the aluminate
amounts of Na2O into the cubic structure of C3A leads to a symmetry and silicate phases, it sometimes can be necessary to study single phase
reduction from Pa3 (cubic) →P21 3 (cubic II) →Pbca (orthorhombic) → models such as pure C3A polymorphs. Studying their hydration behavior
P21 /a (monoclinic) [10–13]. It has been demonstrated that the reaction is relevant in research efforts directed towards early cement reactivity.
of cubic C3A with water is faster than orthorhombic C3A in the absence Accordingly, the lab synthesis of pure cement phases has been a
of CaSO4. In contrast, the hydration rate of the cubic polymorph is cornerstone of cement research in the past. Traditionally, pure cement
slower compared to orthorhombic and monoclinic polymorphs if CaSO4 clinker phases are prepared via solid state synthesis, often accompanied
is present [13–15]. In commercial Portland cement, calcium sulfate by long sintering times (up to 24 h per cycle) and several burning cycles
sources such as gypsum (CaSO4⋅ 2 H2O), bassanite (CaSO4⋅ 0.5H2O) or (3 to 5). Thus, a synthesis of pure C3A is stretched over days and can be
anhydrite (CaSO4) are interground with cement clinker. The purpose of described as a repeating pattern of a grinding step, followed by me­
the calcium sulfate source is to suppress the initial dissolution of C3A and chanical compaction and a final high-temperature sintering step
thus to prevent very fast setting, also known as flash setting. Further­ [19–22]. The sintering process to form C3A can be accelerated by the
more, the total calcium sulfate amount has to be carefully adjusted in preparation of very densely packed specimen prior to sintering. This
Portland cement to ensure that the sulfate depletion occurs after the heat usually requires substantial mechanical pressures (e.g., > 100 MPa and

* Corresponding author.
E-mail address: torben.gaedt@tum.de (T. Gädt).

https://doi.org/10.1016/j.cement.2023.100064
Received 2 February 2023; Accepted 27 February 2023
Available online 5 March 2023
2666-5492/© 2023 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-
nc-nd/4.0/).
D. Axthammer et al. Cement 12 (2023) 100064

therefore typically only yields small amounts (low gram scale) [23]. In Table 1
2004, Stephan and Wilhelm [24] presented a sol-gel route to pure Raw material mix compositions for all prepared C3A samples of this study.
clinker phases, including C3A. Here, the gel is prepared from a mixture Sample CaO CaCO3 Al2O3 Al2O3 Na2O NaNO3 CAC
of Ca(NO3)2 and a commercial alumina sol by adjusting the pH value, lables
followed by a drying and a sintering step. The firing process had to be mol g mol g mol g g
conducted two to three times for 14 h in the case of pure cubic C3A. No C3A_S_0 1.11 111.12 0.37 37.74 - - -
specification on the yield of this sol-gel method was provided by the C3A_S_4.6 1.05 105.17 0.37 37.74 0.08 12.76 -
authors. An alternative sol-gel route to C3A based on an aqueous solu­ C3A_S_6.2 1.01 101.10 0.37 37.74 0.10 16.98 -
C3A_L_0 6.25 565.50 2.08 121.10 - - 125.00
tion of Ca(NO3)2 and Al(NO3)3 which is gelled by a polycondensation of C3A_L_4.6 5.90 530.48 2.08 121.10 0.42 71.40 125.00
the auxiliary components citric acid and ethylene glycol (also known as C3A_L_6.2 5.68 508.18 2.08 121.10 0.56 95.19 125.00
the Pechini method [25]) has been reported by Gaki et al. [26] The green
body is obtained via the esterification during heating and is sintered (at
1000 ◦ C) subsequently. Additionally, Voicu et al. described a the resulting high viscosity suspension in molds to obtain self-supporting
Pechini-type synthesis in ethanol for C3A which only needs one firing cylinders, which were subsequently dried and sintered. Their described
cycle and a temperature of 1350 ◦ C for 1 h. The reported yield is 10 g of method resulted in yields of 400 g C3S per batch in high quality. Georget
the final product [27]. While this method should be scalable, it is based et al. [34] published a related preparation route for C3A, with a long
on a diluted ethanol solution and a molar ratio of citric acid to cation of milling time of 30 h and subsequent drying of the green body at 105 ◦ C
1. Therefore, a kilogram of tricalcium aluminate needs 4.5 L of ethanol for approximately one month.
and 3.09 kg of citric acid (based on the synthetic details provided in the Inspired by the work of Li et al., we report a convenient method for
paper) which is not practical (and expensive) for lab-scale synthetic the preparation of C3A polymorphs on a scale of 500 g per batch. By the
procedures. Another synthesis route towards C3A is based on mixtures of preparation of tall cylindrical self-supporting green bodies, we maxi­
the nitrate salts of calcium and aluminium and their combustion with mize the C3A yield per crucible compared to the conventional powder
urea at elevated temperatures, also referred to as self-propagation compaction methods. Crucially, commercially available calcium
combustion. Here, mixtures with urea only and urea/alanine fuels aluminate cement is used both as a binder to obtain green strength
have been reported. In both cases, a short and relatively cylinders and as a co-reactant supplying calcium and aluminium ions.
low-temperature (900-1050 ◦ C) classic sintering step is required after For the synthesis, we decided to minimize the amount of calcium
the combustion to achieve full conversion to C3A. The reported protocols aluminate cement as binder. Hence, we used just enough CAC to achieve
yield about 10 g and increasing the scale by a factor of 100 to achieve sufficient green body strength, while we ensure that the major fraction
kilogram scale yields would include significant safety considerations of raw materials is similar to conventional synthesis routes. This concept
and is therefore not practical [28,29]. can be expanded to the synthesis of other phases by using a reactive
In an article published in 2009, Wesselsky and Jensen [22] sum­ binder material to prepare self-supporting cylinders, that additionally
marized lab-scale synthesis routes for pure and doped C3A. The provide the required elements for the desired phase. Another advantage
described synthesis protocols include a minimum of two sintering steps. of the new binder based process is its simplicity. At least in our lab,
The first step to form CaO from CaCO3, usually directly accompanied by coworkers often perceived mechanical powder compaction of CaCO3
the reaction of CaO with Al2O3 and further steps to increase phase purity and Al2O3 as cumbersome and showed a strong preference for the new
above 99 wt.%. The recipes in the publication are calculated to yield binder based route. The purity of the obtained C3A polymorphs were
100 g of pure phases. In the described method, powders are compacted characterized by X-ray powder diffraction and X-ray fluorescence
into platinum crucibles by pressing the powders with an appropriately spectroscopy and compared to C3A polymorphs obtained using the
shaped tool or by employing pellets obtained with a mechanical press. classic synthesis route. Their reactivity was investigated using in-situ
The yields are limited by the size of available crucibles, the space of the isothermal heat flow calorimetry. Additionally, we report slump flow
high temperature furnace and the volume of the milling devices. Addi­ experiments on C3A pastes as an example for a method which requires
tionally, the authors suggested two to five firing cycles to achieve material amounts in excess of 10 g C3A per test.
products in high purity over 99 wt.%, resulting in additional preparation
efforts. 2. Materials and methods
In summary, there is a lack of laboratory methods to prepare pure
C3A polymorphs in amounts significantly larger than 100 g, due to the 2.1. Materials
limited amount of raw material per platinum crucible. Such amounts are
desirable because methods like strength tests according to DIN EN All applied chemicals (with purity) in this study are commercially
196-1:2016-11 require 450 g ± 2 g of the binder (Portland cement in available and were used without further purification: CaCO3 (99.3%),
case of the DIN test) [30]. In contrast to experiments with synthetic Al2O3 (99.7%), NaNO3 (99.5%), calcium aluminate cement (27% CaO
polyphase clinker, it can be desirable to conduct measurements with and 72% Al2O3) and quartz powder (SiO2 98.0%).
cementitious model systems here (clinker phase mixed with inert filler
or additives, e.g., supplementary cementitious materials). If such a 2.2. Sample preparation
model mixture with a typical amount of 4-10 wt.% C3A has to be tested,
18-45 g of anhydrous C3A per test would be consumed. This results in Raw material mixes for all prepared C3A samples of this study are
only two to five tests per C3A batch of the literature known synthesis. summarized in Table 1. Samples prepared with the small scale synthesis
Additionally, larger batch sizes ensure that larger experimental pro­ were named as C3A_S_0, C3A_S_4.6 and C3A_S_6.2. The number corre­
grams can be done with the same C3A batch, avoiding differences be­ sponds to the theoretical Na2O content in the raw material mix of 0, 4.6
tween different synthesis batches [13,31,32]. and 6.2 wt.% respectively. The same designation applies to the poly­
In the pursuit of obtaining larger yields from pure phase lab syn­ morphs prepared by the large scale method, which are C3A_L_0,
thesis, Li et al. [33] recently published a method to produce C3S on the C3A_L_4.6 and C3A_L_6.2.
kilogram scale. The preparation of stable green bodies directly from
CaCO3 and SiO2 without the need for a polymer binder or a classic 2.2.1. Small scale synthesis
sol-gel process facilitates an increase in the yield of C3S, which can be Small scale syntheses for the different C3A polymorphs were per­
obtained in one sintering step. For this purpose, the authors milled an formed according to Wesselsky and Jensen [22]. Stoichiometric mix­
aqueous suspension of fumed silica with CaCO3 in a ball mill and poured tures of CaCO3, Al2O3 with and without NaNO3 (according to Table 1)

2
D. Axthammer et al. Cement 12 (2023) 100064

Table 2
Detailed procedure for the large-scale synthesis of cubic C3A. The sintering
temperature is reduced to 1300 ◦ C for the sodium substituted polymorphs.
Step Procedure Duration

Homogenization Tumbling 2h
Milling 1h
Addition of water Molding and Hydration 14 h
Drying 105 ◦ C 4h
Sintering RT to 1000 ◦ C 6h
1000 ◦ C 4h
1000 ◦ C to 1400 ◦ C 4h
1400 ◦ C 6h
Preparation for Sintering Milling and Compacting 1.5 h
Sintering RT to 1400 ◦ C 8h
1400 ◦ C 6h
Final preparation Milling and Sieving 1.5 h

Fig. 2. Cylinders during the large scale synthesis of cubic C3A. Left: Green
cylinder after hydration with a height of 10 cm and right: Sintered cylinder with
a height of 7.5 cm.

polymorphs in order to reduce the loss of volatile Na2O. Additionally, a


detailed temperature profile of the synthesis process is displayed in
Fig. 1.
Table 1 shows the mixes yielding 500 g of pure C3A after sintering.
According to the desired composition, CaCO3, Al2O3, CAC with and
without NaNO3 were weighed together and tumbled in an overhead
shaker (Heidolph REAX 20) at maximum rotational speed for 2 h.
Afterwards, the mixture was ground in a rotary ball mill (AAM-WA 350,
AAM Mahltechnik, Germany) with 3 kg of ZrO2 grinding balls (diameter
Fig. 1. Synthesis steps and temperature profile for the large scale synthesis of
= 12.5 mm) at 70 rpm for 1 h. The homogenized powder was then mixed
cubic C3A. Sintering temperature for the sodium doped polymorphs is reduced
with water (w/s = 0.5) for 2 min at 600 rpm using an overhead stirrer
to 1300 ◦ C in order to minimize evaporation of Na2O.
equipped with a propeller paddle. The fresh pastes (which had a
homogeneous and pasty consistency) were immediately filled in silicone
were ground in a planetary ball mill (Pulverisette 6, Fritsch) for 15 min
molds and were hydrated at room temperature overnight (typically
at 300 rpm. The sodium amounts were chosen to slightly exceed the
14 h). These timeframes can be easily adopted in typical laboratory
values for the required substitution of orthorhombic and monoclinic C3A
workflows. After hydration, the green cylinders were first dried at
mentioned in the literature to overcome potential losses due to evapo­
105 ◦ C for 4 h and then placed in platinum crucibles. For the sodium
ration of volatile Na2O [11]. The powders were loaded into platinum
containing batches, the crucibles were covered with additional inverted
crucibles in four portions and were intermediately compacted by
crucibles in order to reduce evaporation of Na2O. The sintering protocol
pressing the powders into the crucibles with a spatula and a pistil. The
was set to be equal to the small scale sintering protocol, with a decar­
crucibles were covered with platinum-rhodium lids. Sintering was
bonization step at 1000 ◦ C followed by calcination for 6 h at maximum
conducted by a calcination step at 1000 ◦ C followed by firing the sam­
temperature. A representative photograph of the green cylinder before
ples at 1400 ◦ C for the cubic polymorph and at 1300 ◦ C for the sodium
the sintering and the sintered cylinder is shown in Fig. 2. After the first
containing samples, in each case for 6 h. An intermediate grinding step
sintering step, the samples were ground and mechanically compacted
was conducted prior to the second sintering step to increase reactivity.
into the crucibles. Again, a second firing step at 1400 or 1300 ◦ C for 8 h
Again, the ground powders were compacted into the crucibles and sin­
was conducted. Samples were quenched with pressurized air and ground
tered at 1400 or 1300 ◦ C for 8 h. After the indicated duration, samples
in the rotary ball mill. The powders were sieved with a mesh size of 90
were removed from the oven, were quenched with pressurized air and
μm. This procedure resulted in about 500 g per C3A batch in three days.
ground for 15 min at 300 rpm. The powders were sieved with a mesh size
The obtained samples were stored in a desiccator under vacuum over
of 90 μm. This preparation method resulted in yields of approximately
silica gel as drying agent until further usage. Additionally, for one batch
100 g per batch C3A.
of the cubic polymorph (C3A_L_DS: large scale cubic C3A directly sin­
tered (DS); raw mixture analogous to C3A_L_0), the second sintering step
2.2.2. Large scale synthesis
was applied to the sintered cylinders without intermediate grinding.
Large scale laboratory syntheses of different C3A polymorphs started
by producing stable green cylinders of the respective raw materials.
Calcium aluminate cement (CAC) was added to the reagents to prepare 2.3. Sample characterization
cylinders with appropriate strength to be self-supporting and to increase
the amount of raw material per crucible. The oxide composition of the X-ray diffraction patterns of all samples were recorded on a Bruker
CAC was specified to 27% CaO and 72% Al2O3 by the manufacturer and D8 Advance diffractometer in Bragg-Brentano θ-2θ geometry equipped
was confirmed by XRF measurements (27.3% CaO, 72.1% Al2O3, 0.3% with a VÅNTEC-1 LPSD detector (12◦ 2θ detector opening). The goni­
SiO2, 0.1% Fe2O3). The amount of calcium and aluminium in the cal­ ometer radius was 217.5 mm and the X-Ray source emitted CuKα1,2
cium aluminate cement is accounted for and subtracted from the mix of radiation (1.5406 Å). Powder XRD measurements were conducted at
CaCO3 and Al2O3 according to the stoichiometric ratio required to 32 kV and 40 mA. The divergence slit was fixed to 0.2◦ and primary and
obtain C3A. An exemplary step-by-step procedure for the synthesis of secondary soller slits of 2.5◦ were used. The diffraction data was
cubic C3A is shown in Table 2. The maximum sintering temperature was collected between 7◦ and 70◦ 2θ with a step size of 0.016◦ 2θ and a
1400 ◦ C for cubic C3A and was reduced to 1300 ◦ C for the other counting time of 0.5 s per step, which resulted in a total scan time of
40 min. The samples were prepared by the front loading technique and

3
D. Axthammer et al. Cement 12 (2023) 100064

Table 3 scale C3A samples exhibit a finer particle size distribution compared to
ICSD collection codes used for the Rietveld analysis of the synthesized C3A the large scale preparations. Additionally, Blaine values are noticeably
samples. reduced for the large scale samples (from small scale: 5800-6300 cm2/g
Phase ICSD Code Reference to large scale: 4000-4700 cm2/g).
C3A cubic 1841 [36]
A comparison of PSD of the two prepared cubic C3A polymorphs is
C3A ortho. 1880 [37] displayed in Fig. 3. The small scale samples were ground using the
C3A mono. 100221 [38] planetary ball mill, producing finer particles due to a larger energy input
C12A7 (Mayenite) 261586 [39] compared to the large scale samples ground with the rotary ball mill.
Lime 75786 [40]
Nevertheless, a trimodal particle size distribution is observed for the
small scale samples. The application of the rotary ball mill in the larger
scale synthesis leads to a more uniform PSD and slightly larger values for
Table 4 Dv50 (C3A_S_0: 3.01μm and C3A_L_0: 8.82 μm) and Dv90 (C3A_S_0: 23.17
Particle size distributions and Blaine specific surface of synthesized C3A samples μm and C3A_L_0: 33.99μm). According to our findings, the use of a low
via small and large scale laboratory protocols. energy rotary ball mill (used for the large scale samples) with slightly
Sample Dv10 Dv50 Dv90 Blaine longer grinding time appropriate for the preparation of C3A.
μm μm μm cm2/g

C3A_S_0 0.27 3.01 23.17 5852


C3A_S_4.6 0.26 2.97 28.40 6099
2.5. Isothermal calorimetry
C3A_S_6.2 0.27 3.21 22.92 6276
C3A_L_0 0.70 8.82 33.99 4004 Measurements of reaction heat were conducted on an eight-channel
C3A_L_4.6 0.59 6.02 25.26 4660 isothermal heat flow calorimeter (TAM Air, TA Instruments, USA) at
C3A_L_6.2 0.55 6.35 28.90 4444
20 ◦ C. Quartz sand was used as reference material, with an amount
(11.4 g) matching the heat capacity of the measured samples calculated
according to Wadsö [41]. Commercial in-situ mixing devices were used
(Bohr-O-Mir 4000, Version 4, Technisch Zeichnen Grassl, Germany) in
combination with reaction vials made of a plastic/graphite composite.
1.0000g ± 0.0005 g anhydrous C3A powder was weighed in the reaction
vials and 2.0000 g ± 0.0005 g deionized water was weighed into two
syringes. The in-situ mixers were placed in the calorimeter and left to
equilibrate overnight. The mixing protocol consisted of 10 s pre-mixing
of the dry powder followed by the injection of water and 60 s continuous
stirring using an electrical motor. Reactions were typically monitored
over 24 h.

2.6. Slump flow

10 g of the respective C3A polymorph powder were weighted in a


Fig. 3. Comparison of the particle size distributions of cubic C3A ground using plastic cup and 30 g of a fine quartz powder (Sikron SF 300) were
the planetary ball mill (small scale synthesis, C3A_S_0) and the rotary ball mill added as inert filler material. The powders were mixed for 10 s at 600
(large scale synthesis, C3A_L_0). rpm with an overhead stirrer. Afterwards, 40 g deionized water was
added (w/s = 1), and the mixture was stirred 2 min at 600 rpm using the
were rotated with 30 rpm during the measurements. Rietveld same stirrer. Directly after mixing, the pastes were poured into a brass
refinements and phase quantification were carried out using the Topas- cylinder (diameter = 30 mm, height = 50 mm, according to DIN EN
Academic V7 software [35]. Anhydrous phases with corresponding ICSD 12706:1999-12 [42]) placed on a glass plate. The cylinder was vertically
codes for the evaluation of the purity of the synthesized C3A samples by lifted at 3 min after water addition. The resulting spread was measured
Rietveld analysis are given in Table 3. in three directions and the mean value was calculated.
The oxide compositions were investigated by X-ray fluorescence
spectroscopy. The necessary tablets were prepared by homogenizing the
2.7. Compressive strength
C3A powders or calcium aluminate cement with lithium tetraborate
(Merck). The mixtures were melted and several granules of lithium io­
We tested the compressive strength of one exemplary raw mixture
dide (Merck) were added. These melted mixtures were poured into hot
paste to ensure that cylinders developed appropriate strength overnight.
platinum molds and cooled with pressurized air afterwards. The
Therefore, small scale compressive strength tests were conducted after
resulting tablets were analyzed using an AXIOS XRF spectrometer
14 h (time of demolding of cylinders) using the hydrated raw mixture
(Malvern Panalytical GmbH, Germany).
paste of C3A_L_0, similar to DIN EN 12390-3:2019-10 [43]. Therefore,
60 g of the pre-mixed powder was weighted in a plastic cup, 30 g
2.4. Particle size distributions of deionized water was added (w/s = 0.5), and the mixture was stirred
2 min at 600 rpm using an overhead stirrer. Directly after mixing, the
Particle size distributions (PSD) of the synthesized and milled C3A pastes were poured into cubic molds (15 x 15 x 15 mm) and were hy­
samples were determined through laser diffraction analysis using a drated at room temperature overnight, analogous to the hydration of the
CILAS 1064 Particle Size Analyzer. Prior to the measurements, the cylinders. The resulting cubes were demolded after 14 h, the compres­
anhydrous powders were suspended in isopropanol and pretreated for sive strength was measured with a loading speed of 0.6 MPa/s and the
10 min in an ultrasonic bath to avoid agglomeration of particles. mean value as well as the standard deviation were calculated from
PSD of the synthesized C3A samples are given in Table 4. The small 3 measurements.

4
D. Axthammer et al. Cement 12 (2023) 100064

Fig. 4. Schematic illustration of the main steps during the large scale synthesis of the C3A polymorphs. * A direct second burning step applied to the sintered
cylinders for cubic C3A resulted in a high purity product without intermediate grinding (C3A_L_DS).

3. Results and discussion

3.1. Synthesis method

How should an efficient synthesis of C3A polymorphs be designed?


Criteria which should be met are: a) it should have as few steps as
possible, b) only use cost-efficient, commercially available raw mate­
rials, c) be as atom-efficient as possible, d) be benign in terms of toxi­
cology and safety, e) enable the use of anisotropic green bodies, e.g.,
long cylinders, to minimize the need for expensive platinum as crucible
material. We believe that these criteria are met by preparing cylindrical
green bodies using commercial calcium aluminate cement as a binder
Fig. 5. Powder XRD pattern after first sintering of the large scale sample of
and co-reactant for the synthesis of C3A from commercial CaCO3 and
cubic C3A (C3A_L_0). Contributions shown for cubic C3A (gray), mayenite (blue)
Al2O3 sources. In the context of this study, the exact phase composition
and lime (orange). (For interpretation of the references to colour in this figure
of the employed CAC was not determined. Nevertheless, it is known that legend, the reader is referred to the web version of this article.)
commercial calcium aluminate cements with comparable oxide com­
positions (i.e., 70% Al2O3 and 30% CaO) are mainly composed of the
Here, the pre-mixed reactants including the calcium aluminate cement
phases CA and CA2 [44,45], and consist of approximately 60 wt.% CA,
were mixed with water and hydration reaction was typically carried out
33-39 wt.% CA2 and 1-6 wt.% Al2O3 [46,47]. The hydration behavior of
overnight (the final setting time of CAC according to the manufacturer is
these phases is a rather complex process and a detailed description is out
6 h). After overnight hydration for 14 h, the cylinders had developed
of the scope of this paper and has been described in detail elsewhere[44,
sufficient strength to be easily demolded. The time of 14 h was chosen
46,48,49]. In summary, the main phase CA reacts with water at room
based on practical workflow considerations. The samples are prepared
temperature to initially form C2AH8, accompanied by the formation of
and filled into molds in the late afternoon, and demolding and drying
CAH10 and AH3 [50]. The phases C2AH8 and CAH10 are metastable and
commences the next morning. In an additional experiment, we evalu­
ultimately convert to thermodynamically stable phase C3AH6. In our
ated the compressive strength of the material using small cubic molds
case, the green strength of the cylinders is most likely due to CAH phases
(15 x 15 x 15 mm). We found a compressive strength of the raw mixture
like C2AH8, which form in the early phase of hydration from CA and
of 0.40 MPa (σ = 0.04 MPa) after hydration overnight for 14 h. After
water. In contrast to CA, the reaction of CA2 with water is significantly
demolding of the cylinders, the excess water was removed by
slower and probably does not contribute to the strength of the cylinders
oven-drying (4 h at 105 ◦ C) to avoid damage due to fast water evapo­
in our case (demolding occurred after 14 h) [51]. The determination of
ration during the early stage of the high-temperature sintering step. The
the exact phase composition of the binder matrix was not in the focus of
following sintering steps were similar to reported solid state preparation
this study. Instead, the main objective is the preparation of large scale
methods, (i.e., 6 h at 1300 or 1400 ◦ C respectively) [31,33,52]. For cubic
self-supporting green bodies using a binder which also supplies the
C3A, we experimentally confirmed that the intermediate grinding step
necessary ions for the final phase C3A and therefore is atom efficient. In
between the first and second sintering step can also be skipped, i.e., the
contrast to other preparation methods producing green bodies
sintered cylinder from the first sintering is simply subjected to a second
(e.g., sol-gel or Pechini methods [24,26,27]), the advantage of the new
sintering step. It can be assumed that this simplification also applies for
method resulted not only from the enlarged sample volume, but also by
the other polymorphs. This has not been experimentally validated
providing calcium and aluminium by the binder material itself. Addi­
though. It is important to note, that all hydration experiments were done
tionally, on the contrary to combustion methods [28,29], no extra safety
using C3A polymorphs which had been synthesized by including the
considerations (on top of the necessary precautions for handling very
intermediate milling step depicted in Fig. 4. In general, with the use of
high temperature materials) are necessary. A schematic illustration of
calcium aluminate cement as binder, the method only includes one
the major steps for the large scale synthesis of C3A is shown in Fig. 4.

Table 5
Results of Rietveld analysis carried out with Topas-Academic V7 of synthesized C3A samples via small and large scale laboratory protocols and the detected amount of
Na2O with XRF.
Scale Sample C3A c. C3A o. C3A m. C12A7 Lime Rwp Rp Na2OXRF
wt.% wt.% wt.% wt.% wt.% % % wt.%

Small C3A_S_0 99.4 - - 0.3 0.3 6.7 5.2 0.1


C3A_S_4.6 - 99.2 - 0.5 0.3 5.7 4.5 3.8
C3A_S_6.2 - - 99.5 0.4 0.1 6.2 4.7 4.5
Large C3A_L_0 99.9 - - 0.0 0.0 10.2 7.7 -
C3A_L_4.6 - 97.9 - 1.6 0.5 6.9 5.2 3.1
C3A_L_6.2 - - 99.1 0.2 0.7 7.4 5.5 4.1
C3A_L_DS 99.7 - - 0.2 0.1 9.0 6.9 -

5
D. Axthammer et al. Cement 12 (2023) 100064

amounts of impurities of mayenite (C12A7) and lime. One exemplary


XRD pattern after the first sintering step (C3A_L_0: 2.7 wt.% mayenite
and 2.5 wt.% lime) is shown in Fig. 5. Consequently, two burning steps
were chosen for all six syntheses, with intermediate grinding and com­
pacting applied to all samples. Essentially, both synthetic methods
resulted in C3A polymorphs with a high purity (> 99 wt.%) and with low
amounts of residual lime and mayenite after the second burning step.
Only one batch (C3A_L_4.6) contains a remaining impurity of mayenite
(1.6 wt.%).
As an example, the XRD pattern of the cubic large scale sample
Fig. 6. Powder XRD pattern of the synthesized large scale sample of cubic C3A
(C3A_L_0) after the second sintering is shown in Fig. 6. The residuals
(C3A_L_0). Calculated fit by Rietveld analysis (orange) is displayed together confirm the good quality of the refinement. Also, the agreement indices
with the background (blue) and residue (gray). (For interpretation of the ref­ presented in Table 5 (Rwp: 5.7-10.2 % and Rp: 4.5-7.7 %) for all Rietveld
erences to colour in this figure legend, the reader is referred to the web version analyses are low, again confirming the sufficient fit quality and phase
of this article.) purity. Hence, with both methods, high purity C3A samples were ob­
tained. Additionally, the cubic C3A sample sintered for a second time
without intermediate grinding (C3A_L_DS) also exhibit high purity with
Table 6 99.7 wt.% (C3A cubic).
Results of the XRF analysis of synthesized C3A samples via small and large scale To investigate the amount of sodium oxide incorporated in the C3A
laboratory protocols.
crystal structures, XRF analyses were conducted. The results of these
Sample CaO Al2O3 Na2O ΔNa2O SiO2 measurements are given in Table 6. Although the platinum crucibles
wt.% wt.% wt.% % wt.% were covered with lids (small scale) or inverted crucibles (large scale)
C3A_S_0 62.5 35.5 0.1 - 0.2 during the sintering, the evaporation of Na2O is observable for all so­
C3A_S_4.6 58.2 35.5 3.8 17.4 0.2 dium containing polymorphs. The relative loss between the theoretically
C3A_S_6.2 57.1 35.5 4.5 27.4 0.2
expected sodium oxide content and the detected amount is indicated as
C3A_L_0 62.1 35.5 - - 0.3
C3A_L_4.6 58.3 35.7 3.1 32.6 0.3 ΔNa2O in Table 6. Due to a larger total surface area available for evap­
C3A_L_6.2 56.6 35.8 4.1 33.9 0.3 oration of the large cylinders, the loss of Na2O is more prominent
compared to the small scale reactions. A small amount of SiO2 is
detected in all six samples, most likely caused by minor impurities in the
simple mixing step with water to produce large green cylinders with no applied raw materials. Detected amounts of Na2O in the substituted
further addition of other materials and is therefore easy to conduct. orthorhombic and monoclinic polymorphs are slightly below the values
given in the literature for these phases (about 0.5 wt.%) [11,22].
3.2. Quality of C3A samples Nevertheless, Rietveld analysis (Table 5) confirmed the purity of the
products. The deviations may occur from uncertainties in the deter­
Results of the Rietveld refinements after two sintering cycles with mined amounts by XRF analysis.
intermediate grinding and compacting for all prepared samples are
given in Table 5. After one sintering, all samples included non-negligible

Fig. 7. Cumulative heat (A) over 24 h and heat flow (B) during the first 60 min of the C3A polymorphs prepared via small scale lab syntheses (top) in comparison
with cumulative heat (C) over 24 h and heat flow (D) during the first 60 min of the large scale C3A samples (bottom).

6
D. Axthammer et al. Cement 12 (2023) 100064

Table 7
Summary of hydration characteristics of synthesized C3A samples measured
with isothermal calorimetry.
Sample Peak height Peak time Heat J/(g C3A)

W/(g C3A) min 1h 3h 12 h 24 h

C3A_S_0 0.62 5.7 651 787 909 922


C3A_S_4.6 0.36 5.0 413 545 644 651
C3A_S_6.2 0.18 5.6 223 293 353 375
C3A_L_0 0.47 5.7 615 769 896 930
C3A_L_4.6 0.35 4.5 463 681 793 802
C3A_L_6.2 0.19 4.9 280 422 513 527

Fig. 9. The effect of sodium substitution on cumulative heat after 24 h for all
prepared C3A samples.

of sodium incorporated in the C3A structure can explain these differ­


ences. Fig. 8 compares cumulative heat curves for all six prepared C3A
samples over 24 h. Both unsubstituted cubic C3A samples resulted in
almost equal heat values after 24 h, with only minor deviation in the
early reaction period already explained before. We did not determine
the degree of hydration experimentally. But for both cubic polymorphs
(i.e., C3A_L_0 and C3A_S_0), we assume a degree of hydration of at least
90%, based on the calculation of the theoretical heat of complete hy­
dration of C3A to form katoite (− 968 J/g, for comparison, the expected
Fig. 8. Comparison of cumulative heat of small and large scale syntheses over heat of full conversion of C3A into C4AH13 and C2AH7.5 is − 1034 J/g)
24 h of hydration. using the CEMDATA18 database and the experimental heat after 24 h of
922 J/g and 930 J/g [54]. With an increase in Na2O substitution, the
hydration reaction of C3A is more and more suppressed and curves
3.3. Hydration behavior follow the trends according to the results in XRF analyses shown in
Table 6. As described in the previous section, the large scale syntheses
The heat flow curves representing the first 60 min of hydration of the resulted in an increased loss of Na2O. Consequently, cumulative heat
C3A samples are represented in Fig. 7. Analogous to previous studies, curves demonstrate higher values in measured heat compared to the
cubic C3A has the fastest rate of hydration in pure water, with a small scale samples. The sodium substitution in the prepared C3A sam­
maximum of 0.62 W/(g C3A) for the material obtained via small scale ples strongly affects their hydration behavior also displayed in Fig. 9.
protocol C3A_S_0 (see Fig. 7 B). With increased Na2O-content, the initial With the Na2O substitution exceeding 3 wt.%, an approximately linear
reaction is slowed down, resulting in lower measured thermal powers of dependency between the sodium content and measured cumulative heat
0.36 and 0.18 W/(g C3A) respectively. This effect is often explained by a after 24 h can be recognized. In contrast to the findings of Stephan and
change in dissolution kinetics of C3A due to the Na2O substitution in the Wistuba (2006) [15], significant reductions of heat values after 24 h are
C3A crystal structure [13–15,53]. The cumulative heat of the small scale detected when Na2O is present in the C3A crystal structure. To further
samples is also highly influenced by the incorporated Na2O in the crystal evaluate the dependency between measured heat and Na2O, additional
structure of C3A (Fig. 7 A). After 24 h, heat is reduced from 922 to 375 experiments would be needed to close the gap of substituted amount
J/(g C3A) if cubic C3A is compared to the polymorph (C3A_S_6.2) with between 0 and 3.1 wt.%.
the highest sodium content. Our findings on hydration kinetics of cubic The key values for the hydration characteristics of the prepared C3A
and orthorhombic C3A in pure water are in agreement with previous samples are summarized in Table 7. As discussed before in this section,
studies [14]. the rising sodium substitution clearly influences the maximum heat flow
The heat flow of the first hour of hydration of large scale C3A samples in the early phase of hydration and decrease the peak height by over
is shown in Fig. 7 D. The shapes of the curves are similar to the small 50% (from unsubsituted C3A to the highest subsitution tested). Addi­
scale samples. Furthermore, the samples follow the same trend, i.e., the tionally, this effect is accompanied by a decrease of all measured cu­
cubic polymorph shows the fastest reaction while an increasing Na2O mulative heat values for the substituted polymorphs. In summary, large
substitution slows down the initial reaction. The substituted polymorphs scale samples showed analogous hydration behavior (depending on so­
exhibit comparable values to the small scale samples (C3A_L_4.6: 0.35 dium doping and PSD) as conventionally synthesized C3A.
W/(g C3A), C3A_L_6.2: 0.19 W/(g C3A), also see Table 7). In contrast, the
reduced heat flow maximum of the large scale cubic sample with a
3.4. Storage stability
measured thermal power of 0.47 W/(g C3A) compared to 0.62 W/(g
C3A) for the small scale sample can be explained by its coarser particles
In the previous section it was shown that large scale syntheses of the
according to the determined PSD (see Table 4, Dv90 (C3A_S_0): 23.17 and
C3A polymorphs resulted in samples with comparable reactivities and
Dv90 (C3A_L_0): 33.99 μm).
analogous trends compared to conventional small scale syntheses.
Additionally, the cumulative heat curves of the three synthesized
Importantly, long-term studies with such materials require sufficiently
large scale samples are displayed in Fig. 7 C. The data again confirms
long storage stability. Dubina et al. [55] found that C3A polymorphs are
that the increasing sodium substitution resulted in a reduction of the
subject to pre-hydration when relative humidity exceeded 55%. Hence, it
reaction heat after 24 h (C3A_L_0: 930, C3A_L_4.6: 802, C3A_L_6.2: 527 J/
is essential to avoid contact of the samples with humid air. To evaluate
(g C3A)). However, the observed decrease of the heat after 24 h is not as
storage stability, a large scale batch of cubic C3A (C3A_L_0) was stored in a
prominent as for the small scale samples. The higher loss of Na2O during
desiccator over silica gel and under vacuum (100 mbar) to prevent pre­
preparation of the large scale samples and the resulting smaller amount
hydration. The reactivity of the stored samples was determined by

7
D. Axthammer et al. Cement 12 (2023) 100064

three C3A polymorphs prepared at large scale. In particular, we studied a


mixture of 25 wt.% C3A and 75 wt.% quartz powder in order to reduce
the initial heat development of the samples upon mixing with water.
Pastes were prepared with a water-solid ratio of 1 (water to C3A ratio =
4). The dependency of slump flow on the sodium amount in the C3A
structure as well as photographs of the resulting spread are displayed in
Fig. 11. The slump flow increased from cubic (B, 6.9 cm) to ortho­
rhombic (C, 7.4 cm) and to monoclinic (D, 8.0 cm) C3A. The higher
reactivity of the cubic polymorph in the absence of calcium sulfate leads
to a faster dissolution followed by a faster precipitation of calcium
aluminate hydrates. Analogous to the so-called flash-set, the faster for­
mation of these AFm phases resulted in a rapid stiffening of the paste of
cubic C3A compared to the sodium substituted samples and therefore a
smaller spread. With the described slump flow tests, we were able to
Fig. 10. Cumulative heats after 1 h, 12 h and 24 h during hydration of cubic
C3A (C3A_L_0) at different sample ages. show that material-consuming techniques are accessible with the large
scale synthesis. That opens up opportunities for new methods to inves­
isothermal calorimetry at 3, 9 and 12 months of storage and compared to tigate further properties of C3A and its polymorphs.
the freshly synthesized sample. Fig. 10 displays the changes in cumulative
heats after 1, 12 and 24 h respectively. A small decrease of the initial heat 4. Conclusions
after 1 h is found over the tested period. In contrast, values for cumulative
heats after 12 and 24 h remain unaltered over 12 months. Dubina et al. We demonstrated a simple route for the preparation of C3A poly­
described the formation of C-A-H when C3A underwent prehydration morphs on a 0.5 kg laboratory scale. As a key component, the protocol is
[56]. These formed phases should reduce the overall heat measured based on the use of commercial calcium aluminate cement as binder and
during hydration of the stored sample, due to a smaller available amount reactant. The binding properties of CAC allow the preparation of green
of reactive C3A. Although this effect in the early period of hydration can body cylinders, which increase the amount of material per burning cycle
be observed, the reaction heat after 24 h is unchanged. The small decrease that fits on the platinum crucibles. Samples produced via this novel
in initial heat may therefore not be caused by a formation of C-A-H but synthesis protocol showed similar hydration behavior compared to
may be explained by agglomeration of particles or surface passivation samples prepared with the classic lab synthesis based on mechanical
during the period of storage. Consequently, with careful storage, it is powder compaction prior to sintering. A key difference between the
possible to retain reactivity of the prepared C3A samples for at least 1 protocols is the increased evaporation of Na2O during sintering of larger
year. green bodies, which has to be accounted for. Furthermore, under
appropriate storage conditions (i.e., no humidity and carbon dioxide),
the reactivity of the large scale samples remained stable over 12 months.
3.5. Slump flow test A further process simplification is achieved by using the direct synthesis
route, i.e., 2 sintering cycles without intermediate milling which also
To demonstrate a test method which is normally not reported for resulted in high purity cubic C3A, and likely can be applied to other
pure phase investigations (possibly due to limited available amounts), polymorphs. Furthermore, the presented protocol is expected to also
we conducted slump flow tests after 3 min of water addition with all

Fig. 11. A: Dependency between slump flow and sodium oxide content of pastes with large scale C3A polymorphs in a mixture of 25 wt.% C3A and 75 wt.% quartz
powder and a water-solid ratio of 1. B: paste with cubic C3A (C3A_L_0), C: paste with orthorhombic C3A (C3A_L_4.6) and D: paste with monoclinic C3A (C3A_L_6.2).
The cylinder had a diameter of 30 mm and a height of 50 mm.

8
D. Axthammer et al. Cement 12 (2023) 100064

work with other reactive calcium aluminate binders, such as conven­ [16] W. Kurdowski, Cement Hydration, Springer Netherlands, Dordrecht, 2014,
pp. 205–277, https://doi.org/10.1007/978-94-007-7945-7_4.
tionally synthesized C3A instead of commercial calcium aluminate
[17] A. Quennoz, K.L. Scrivener, Interactions between alite and C3a–gypsum hydrations
cement. Finally, we presented results of a material-demanding tech­ in model cements, Cem. Concr. Res. 44 (2013) 46–54, https://doi.org/10.1016/j.
nique such as mini-slump tests as an example for similar testing cemconres.2012.10.018.
protocols. [18] S.T. Bergold, F. Goetz-Neunhoeffer, J. Neubauer, Interaction of silicate and
aluminate reaction in a synthetic cement system: implications for the process of
alite hydration, Cem. Concr. Res. 93 (2017) 32–44, https://doi.org/10.1016/j.
Declaration of competing interest cemconres.2016.12.006.
[19] F.A. Steele, W.P. Davey, The crystal structure of tricalcium aluminate, J. Am.
Chem. Soc. 51 (8) (1929) 2283–2293, https://doi.org/10.1021/ja01383a001.
Torben Gaedt reports financial support was provided by BASF Con­ [20] D. Stephan, H. Maleki, D. Knöfel, B. Eber, R. Härdtl, Influence of Cr, Ni, and Zn on
struction Additives GmbH. the properties of pure clinker phases: part II. C3A and C4AF, Cem. Concr. Res. 29
(5) (1999) 651–657, https://doi.org/10.1016/S0008-8846(99)00008-3.
[21] J.E. Bailey, C.J. Hampson, J. Bensted, P.B. Hirsch, J.D. Birchall, D.D. Double,
Declaration of Competing Interests A. Kelly, G.K. Moir, C.D. Pomeroy, The microstructure and chemistry of tricalcium
aluminate hydration, Philos. Trans. R. Soc. Lond. Ser. Math. Phys. Sci. 310 (1511)
(1983) 105–111, https://doi.org/10.1098/rsta.1983.0070.
The authors declare that they have no known competing financial [22] A. Wesselsky, O.M. Jensen, Synthesis of pure portland cement phases, Cem. Concr.
interests or personal relationships that could have appeared to influence Res. 39 (11) (2009) 973–980, https://doi.org/10.1016/j.cemconres.2009.07.013.
the work reported in this paper. [23] B.M. Mohamed, J.H. Sharp, Kinetics and mechanism of formation of tricalcium
aluminate, Ca3Al2O6, Thermochim. Acta 388 (1) (2002) 105–114, https://doi.org/
10.1016/S0040-6031(02)00035-7.
Data availability [24] D. Stephan, P. Wilhelm, Synthesis of pure cementitious phases by Sol-Gel process
as precursor, Z. Anorg. Allg. Chem. 630 (10) (2004) 1477–1483, https://doi.org/
10.1002/zaac.200400090.
Data will be made available on request. [25] L. Dimesso, Pechini processes: An alternate approach of the sol-gel method,
preparation, properties, and applications, in: L. Klein, M. Aparicio, A. Jitianu
Acknowledgments (Eds.), Handbook of Sol-Gel Science and Technology, Springer International
Publishing, Cham, 2016, pp. 1–22, https://doi.org/10.1007/978-3-319-19454-7_
123-1.
The authors gratefully acknowledge Dr. Xuerun Li for very fruitful [26] A. Gaki, R. Chrysafi, G. Kakali, Chemical synthesis of hydraulic calcium aluminate
discussions. Furthermore T.G. and D.A. acknowledge financial support compounds using the pechini technique, J. Eur. Ceram. Soc. 27 (2) (2007)
1781–1784, https://doi.org/10.1016/j.jeurceramsoc.2006.05.002.
from BASF. [27] G. Voicu, C.D. Ghiţulică, E. Andronescu, Modified pechini synthesis of tricalcium
aluminate powder, Mater. Charact. 73 (2012) 89–95, https://doi.org/10.1016/j.
References matchar.2012.08.002.
[28] R. Ianoş, I. Lazău, C. Păcurariu, P. Barvinschi, Fuel mixture approach for solution
combustion synthesis of Ca3Al2O6 powders, Cem. Concr. Res. 39 (7) (2009)
[1] S. Mantellato, M. Palacios, R.J. Flatt, Relating early hydration, specific surface and
566–572, https://doi.org/10.1016/j.cemconres.2009.03.014.
flow loss of cement pastes, Mater. Struct. 52 (1) (2019) 5, https://doi.org/
[29] A.C. Tas, Chemical preparation of the binary compounds in the calcia-Alumina
10.1617/s11527-018-1304-y.
system by self-Propagating combustion synthesis, J. Am. Ceram. Soc. 81 (11)
[2] H. Uchikawa, K. Ogawa, S. Uchida, Influence of character of clinker on the early
(1998) 2853–2863, https://doi.org/10.1111/j.1151-2916.1998.tb02706.x.
hydration process and rheological property of cement paste, Cem. Concr. Res. 15
[30] DIN EN 196-1:2016-11, Methods of Testing Cement - Part 1: Determination of
(4) (1985) 561–572, https://doi.org/10.1016/0008-8846(85)90053-5.
Strength; German Version EN 196-1:2016, Norm, Beuth Verlag GmbH, Berlin,
[3] F. Winnefeld, A. Zingg, L. Holzer, J. Pakusch, S. Becker, Ettringite-superplasticizer
2016.
interaction and its impact on the ettringite distribution in cement suspensions.
[31] A. Quennoz, K.L. Scrivener, Hydration of C3A–gypsum systems, Cem. Concr. Res.
Proceedings of the 9th ACI International Conference on Superplasticizers and Other
42 (7) (2012) 1032–1041, https://doi.org/10.1016/j.cemconres.2012.04.005.
Chemical Admixtures in Concrete, Sevilla, Spain, 2009, pp. 420.1–420.17.
[32] H. Minard, S. Garrault, L. Regnaud, A. Nonat, Mechanisms and parameters
[4] C. Jakob, D. Jansen, N. Ukrainczyk, E. Koenders, U. Pott, D. Stephan, J. Neubauer,
controlling the tricalcium aluminate reactivity in the presence of gypsum, Cem.
Relating ettringite formation and rheological changes during the initial cement
Concr. Res. 37 (10) (2007) 1418–1426, https://doi.org/10.1016/j.
hydration: a comparative study applying XRD analysis, rheological measurements
cemconres.2007.06.001.
and modeling, Materials (Basel) 12 (18) (2019) 2957, https://doi.org/10.3390/
[33] X. Li, A. Ouzia, K. Scrivener, Laboratory synthesis of C3S on the kilogram scale,
ma12182957.
Cem. Concr. Res. 108 (2018) 201–207, https://doi.org/10.1016/j.
[5] Y. Chen, I. Odler, On the origin of portland cement setting, Cem. Concr. Res. 22 (6)
cemconres.2018.03.019.
(1992) 1130–1140, https://doi.org/10.1016/0008-8846(92)90042-T.
[34] F. Georget, B. Lothenbach, W. Wilson, F. Zunino, K.L. Scrivener, Stability of
[6] R. Ylmén, U. Jäglid, B.-M. Steenari, I. Panas, Early hydration and setting of
hemicarbonate under cement paste-like conditions, Cem. Concr. Res. 153 (2022)
portland cement monitored by IR, SEM and vicat techniques, Cem. Concr. Res. 39
106692, https://doi.org/10.1016/j.cemconres.2021.106692.
(5) (2009) 433–439, https://doi.org/10.1016/j.cemconres.2009.01.017.
[35] A.A. Coelho, TOPAS and TOPAS-Academic: an optimization program integrating
[7] W. Kurdowski, The properties of cement paste, Springer Netherlands, Dordrecht,
computer algebra and crystallographic objects written in C++, J. Appl. Cryst. 51
2014, pp. 279–368, https://doi.org/10.1007/978-94-007-7945-7_5.
(1) (2018) 210–218, https://doi.org/10.1107/S1600576718000183.
[8] P.-C. Aïtcin, 3 - Portland Cement, in: P.-C. Aïtcin, R.J. Flatt (Eds.), Science and
[36] P. Mondal, J.W. Jeffery, The crystal structure of tricalcium aluminate, Ca3Al2O6 31
Technology of Concrete Admixtures, Woodhead Publishing, Cambridge, 2016,
(3) 689–697. doi:10.1107/S0567740875003639">10.1107/S0567740875003639.
pp. 27–51, https://doi.org/10.1016/B978-0-08-100693-1.00003-5.
[37] F. Nishi, Y. Takéuchi, The Al6O18 rings of tetrahedra in the structure of
[9] E. Holt, M. Leivo, Cracking risks associated with early age shrinkage, Cem. Concr.
Ca8.5NaAl6O18 31(4) 1169–1173. doi:10.1107/S0567740875004736.
Compos. 26 (5) (2004) 521–530, https://doi.org/10.1016/S0958-9465(03)00068-
[38] Y. Takéuchi, F. Nishi, I. Maki, Crystal-chemical characterization of the 3Cao⋅Al2O3-
4.
Na2O solid-solution series 152(3–4) 259–307. doi:10.1524/zkri.1980.152.3-4.259.
[10] A.M. Harrisson, 4 - Constitution and specification of portland cement, in: P.
[39] T. Sakakura, K. Tanaka, Y. Takenaka, S. Matsuishi, H. Hosono, S. Kishimoto,
C. Hewlett, M. Liska (Eds.), Lea’s Chemistry of Cement and Concrete (Fifth
Determination of the local structure of a cage with an oxygen ion in Ca12Al14O33,
Edition), Butterworth-Heinemann, Oxford, 2019, pp. 87–155, https://doi.org/
Acta Crystallogr., Sect. B 67 (Pt 3) (2011) 193–204, https://doi.org/10.1107/
10.1016/B978-0-08-100773-0.00004-6.
S0108768111005179.
[11] H. Taylor, Cement chemistry volume 2nd edition, Thomas Telford Publishing,
[40] Q. Huang, O. Chmaissem, J.J. Capponi, C. Chaillout, M. Marezio, J.L. Tholence,
London, 1997, https://doi.org/10.1680/cc.25929.
A. Santoro, Neutron powder diffraction study of the crystal structure of
[12] F.C. Lee, H.M. Banda, F.P. Glasser, Substitution of Na, Fe and Si in tricalcium
HgBa2Ca4Cu5O12+ δ at room temperature and at 10 K, Physica C 227 (1) (1994)
aluminate and the polymorphism of solid solutions, Cem. Concr. Res. 12 (2) (1982)
1–9, https://doi.org/10.1016/0921-4534(94)90349-2.
237–246, https://doi.org/10.1016/0008-8846(82)90010-2.
[41] L. Wadsö, Operational issues in isothermal calorimetry, Cem. Concr. Res. 40 (7)
[13] R.J. Myers, G. Geng, E.D. Rodriguez, P. da Rosa, A.P. Kirchheim, P.J.M. Monteiro,
(2010) 1129–1137, https://doi.org/10.1016/j.cemconres.2010.03.017.
Solution chemistry of cubic and orthorhombic tricalcium aluminate hydration,
[42] DIN EN 12706 :1999-12, Adhesives - Test Methods for Hydraulic Setting Floor
Cem. Concr. Res. 100 (2017) 176–185, https://doi.org/10.1016/j.
Smoothing and/or Levelling Compounds - Determination of Flow Characteristics;
cemconres.2017.06.008.
German Version EN 12706:1999, Norm, Beuth Verlag GmbH, Berlin, 1999.
[14] A.P. Kirchheim, E.D. Rodríguez, R.J. Myers, L.A. Gobbo, P.J.M. Monteiro, D.C.
[43] DIN EN 12390-3:2019-10, testing hardened concrete - Part 3: Compressive strength
C. Dal Molin, R.B. De Souza, M.A. Cincotto, Effect of gypsum on the early hydration
of test specimens; German version EN 12390-3:2019, Norm, Beuth Verlag GmbH,
of cubic and Na-doped orthorhombic tricalcium aluminate, Materials (Basel) 11 (4)
Berlin, 2019.
(2018), 568, https://doi.org/10.3390/ma11040568.
[44] J.H. Ideker, K.L. Scrivener, H. Fryda, B. Touzo, 12 - Calcium Aluminate Cements,
[15] D. Stephan, S. Wistuba, Crystal structure refinement and hydration behaviour of
in: P.C. Hewlett, M. Liska (Eds.), Lea’s Chemistry of Cement and Concrete (Fifth
doped tricalcium aluminate, Cem. Concr. Res. 36 (11) (2006) 2011–2020, https://
doi.org/10.1016/j.cemconres.2006.06.001.

9
D. Axthammer et al. Cement 12 (2023) 100064

Edition), Butterworth-Heinemann, 2019, pp. 537–584, https://doi.org/10.1016/ [51] S.R. Klaus, J. Neubauer, F. Goetz-Neunhoeffer, Hydration kinetics of CA2 and CA –
B978-0-08-100773-0.00012-5. investigations performed on a synthetic calcium aluminate cement, Cem. Concr.
[45] W. Kurdowski, Special cements, Springer Netherlands, Dordrecht, 2014, Res. 43 (2013) 62–69, https://doi.org/10.1016/j.cemconres.2012.09.005.
pp. 603–659, https://doi.org/10.1007/978-94-007-7945-7_9. [52] F.A. Hartmann, J. Plank, Impact of aging on the hydration of tricalcium aluminate
[46] F. Goetz-Neunhoeffer. Modelle Zur Kinetik Der Hydratation von (C3a)/gypsum blends and the effectiveness of retarding admixtures,
Calciumaluminatzement Mit Calciumsulfat Aus Kristallchemischer Und Z. Naturforsch., B: Chem. Sci. 75 (8) (2020) 739–753, https://doi.org/10.1515/
Mineralogischer Sicht, Friedrich-Alexander-Universität Erlangen-Nürnberg, znb-2020-0087.
Erlangen, 2006. Habilitation. [53] A.I. Boikova, A.I. Domansky, V.A. Paramonova, G.P. Stavitskaja, V.
[47] J. Goergens, T. Manninger, F. Goetz-Neunhoeffer, In-situ XRD study of the M. Nikushchenko, The influence of Na2O on the structure and properties of 3 CaO
temperature-dependent early hydration of calcium aluminate cement in a mix with ⋅Al2O3, Cem. Concr. Res. 7 (5) (1977) 483–492, https://doi.org/10.1016/0008-
calcite, Cem. Concr. Res. 136 (2020) 106160, https://doi.org/10.1016/j. 8846(77)90110-7.
cemconres.2020.106160. [54] B. Lothenbach, D.A. Kulik, T. Matschei, M. Balonis, L. Baquerizo, B. Dilnesa, G.
[48] F. Hueller, C. Naber, J. Neubauer, F. Goetz-Neunhoeffer, Impact of initial CA D. Miron, R.J. Myers, Cemdata18: a chemical thermodynamic database for
dissolution on the hydration mechanism of CAC, Cem. Concr. Res. 113 (2018) hydrated portland cements and alkali-activated materials, Cem. Concr. Res. 115
41–54, https://doi.org/10.1016/j.cemconres.2018.06.004. (2019) 472–506, https://doi.org/10.1016/j.cemconres.2018.04.018.
[49] N. Ukrainczyk, Kinetic modeling of calcium aluminate cement hydration, Chem. [55] E. Dubina, J. Plank, L. Black, L. Wadsö, Impact of environmental moisture on C3A
Eng. Sci. 65 (20) (2010) 5605–5614, https://doi.org/10.1016/j.ces.2010.08.012. polymorphs in the absence and presence of CaSO4⋅ 0.5H2O, Adv. Cem. Res. 26 (1)
[50] F. Hueller, J. Neubauer, S. Kaessner, F. Goetz-Neunhoeffer, Hydration of calcium (2014) 29–40, https://doi.org/10.1680/adcr.12.00062.
aluminates at 60◦ c– development paths of C2AHx in dependence on the content of [56] E. Dubina, L. Black, R. Sieber, J. Plank, Interaction of water vapour with anhydrous
free water, J. Am. Ceram. Soc. 102 (7) (2019) 4376–4387, https://doi.org/ cement minerals, Adv. Appl. Ceram. (2013), https://doi.org/10.1179/
10.1111/jace.16314. 174367509X12554402491029.

10

You might also like