Calcium Aluminate Blended
Calcium Aluminate Blended
Calcium Aluminate Blended
Cements Incorporating
Engineered Residues
Jennifer ASTOVEZA
Main supervisors:
Prof. dr. ir. Y. Pontikes
Prof. dr. ir. A. Lecomte
Chairman:
Prof. R. Puers
Reporters:
Prof. S.A.B. Lopez
Dr. ir. V. Baroghel-Bouny
June 2022
École Doctorale Chimie Mécanique Matériaux Physique
Thèse
Directeurs de thèse:
Yiannis Pontikes Professeur des universités
André Lecomte Professeur des universités
Examinateurs:
Romain Trauchessec Maître de conférences (co-directeur)
Ratana Soth Docteur (co-directeur)
Jan Elsen Professeur des universités
Ruben Snellings Docteur
Rapporteures:
Susan Andrea Bernal Lopez Professeure des universités
Véronique Baroghel-Bouny Docteur HDR
Invité:
Robert Puers Professeur des universités (président du jury)
This manuscript is presented in partial fulfilment of the requirements for a dual
PhD from the Faculty of Engineering Science of KU Leuven, and the Institut Jean
Lamour of the Université de Lorraine.
This work was funded by the SOCRATES Project under the European Union
Framework Program for Research and Innovation Horizon 2020 Grant Agreement
No.721385 (EU MSCA-ETN SOCRATES).
Alle rechten voorbehouden. Niets uit deze uitgave mag worden vermenigvuldigd
en/of openbaar gemaakt worden door middel van druk, fotokopie, microfilm,
elektronisch of op welke andere wijze ook zonder voorafgaandelijke schriftelijke
toestemming van de uitgever.
All rights reserved. No part of the publication may be reproduced in any form by
print, photoprint, microfilm, electronic or any other means without written
permission from the publisher.
ISBN 9789464447224
___________________________________________Acknowledgements
Acknowledgements
This work would not have been possible without the contribution of the
following people/organizations:
the team behind the SOCRATES Project for the conception and
funding of this research
Imerys, KU Leuven and Université de Lorraine for the warm
accommodation and support
my thesis advisers, Prof. Yiannis Pontikes, Prof. André Lecomte,
Prof. Romain Trauchessec, and Dr. Ratana Soth for their relentless
supervision, and for always believing in my abilities especially
whenever I am overwhelmed with self-doubt
The members of the jury: Prof. Robert Puers, Prof. Jan Elsen, Dr.
Véronique Baroghel-Bouny, Prof. Susan Andrea Bernal Lopez and
Dr. Ruben Snellings for their constructive review
Dr. Chris Parr and Dr. Christoph Wöhrmeyer of Imerys for offering
me the opportunity to join their team, while providing continuous
support for the completion of my thesis
I would also like to acknowledge several colleagues for their direct
contribution to the experimental works and data analyses, namely:
Dr. Jeroen Soete of KU Leuven for his assistance in tomography and
for performing the analysis on cylindrical volumes in section 3.2.2
Dr. Arne Peys and Christina Siakati of KU Leuven for fitting the
Mössbauer spectra of the raw materials (section 2.2.5) and the
hydrated binary blends (section 3.4.3) using Model 1
Sébastien Diliberto of Institut Jean-Lamour for performing the
Mössbauer spectroscopy and for fitting the spectra in section 3.4.3
using Model 2
Sylvie Migot-Choux and Mélanie Emo of Institut Jean-Lamour for
performing with us the TEM and STEM, respectively
Dr. Tobias Hertel of KU Leuven for his assistance in ATR-FTIR
Dr. Bruno Espinosa of Imerys for the encouragement and for sharing
his insights on the fate of Fe during hydration
my supportive friends and colleagues at Imerys, KU Leuven and
Université de Lorraine for the training and insightful exchanges
Lastly, I would like to express my gratitude to my friends, my second
parents (Michèle & Daniel, Maria & Guido, Zuri & Gonza) and my family
(Nordin, Moises, Everlyn, Glaiza and Maycee). Thank you for choosing to
be part of my life, and for embracing my best and worst.
ASTOVEZA 2022 i
_____________________________________________________ Abstract
Abstract
The pressure to decarbonize the cement industry in light of sustainability
goals has been motivating the search for new types of supplementary
cementitious materials (SCM) in the recent decades. Industrial residues,
which otherwise end up in landfills or find low added-value applications,
are of particular interest in near-zero-waste process schemes.
This thesis explored the potential for valorization of pre-treated residues
as SCM in calcium aluminate cement (CAC)-calcium sulfate hemihydrate
(HH) system. These residues included non-ferrous metallurgy (NFM)
slags, municipal solid waste incinerator (MSWI) bottom ashes, jarosite and
paper-fabrication residues. Using isothermal calorimetry and mechanical
strength test, a screening test was first established to compare their early-
age and the long-term reactivity. The highly-amorphous NFM slags
exhibited superior reactivity showing evidence of long-term (> 28 days)
contribution to the mechanical strength of mortars at 30 wt% replacement
level. However, the slow hydration kinetics in the first days of
hydration could be an issue for CAC systems, given that their applications
greatly rely on their rapid-hardening property. Moreover, the amorphous
nature of the NFM slags and the dominance of iron (Fe) in their
composition, set them apart from traditional SCM . As of this writing,
there are only limited publications discussing the hydration mechanism of
Fe in CAC-based binders.
In order to overcome the slow kinetics of the slag hydration, the influence
of three factors, including: (1) the proportion of sulfates by varying the
CAC/HH ratio; (2) the addition of lime-rich paper residue at 5 wt% as
activator; and (3) the incorporation of 30 wt% PC in ternary CAC-HH-PC
systems, were investigated. For all these formulations, the early hydration
reactions were characterized using isothermal calorimetry on paste
samples. The dimensional stability and mechanical strength were followed
using standard mortars up to 1 year curing period. Furthermore, a
methodology to estimate the slag hydration degree using a non-destructive
technique based on X-ray Computed Tomography (XCT) combined with
volume analysis was developed in this study. Several techniques were used
to follow the phase assemblage evolution, including: X-ray diffraction
(XRD), thermogravimetric analysis (TGA), attenuated total reflection-
Fourier transform infrared spectroscopy (ATR-FTIR). Lastly, transmission
electron microscopy (TEM), scanning transmission electron microscopy
(STEM) and Mössbauer spectroscopy were also performed in order to gain
a better understanding of the fate of Fe, the main component of the slag,
during hydration.
High early strengths (>20 MPa) were obtained for all formulations at
30 wt% slag-binder replacement level. The increased proportion of sulfates
(CAC/HH ratio = 1.6) improved the mechanical strength due to the more
extensive ettringite formation. The addition of PC and paper residue also
improved the strengths at early ages. The slag addition improved the
dimensional stability of mortars by diluting the binder, consequently
mitigating the excessive expansion in the ternary systems. The XCT results
showed that about 50 % of the slag has reacted after more than 180 days
of curing. This was in agreement with the estimations obtained from
Mössbauer spectroscopy, based on the relative amounts of absorption areas
of the newly-formed Fe species during the hydration. While the proportion
dissolution rate
based on the XCT results, the addition of PC and paper residue appeared
to have accelerated it in the first 28 days. At longer term however, the
reaction degree remained relatively comparable to those obtained for the
binary CAC-HH systems. Using the ATR-FTIR spectra, a hypothetical
peak displacement of the Si-O asymmetric stretching band of the slag at
around 930 cm-1 was proposed, supporting this reactivity.
As to the fate of Fe, TEM showed evidence of uptake in hydrated phases
grouped into two compositional clusters when plotted on a Ca-Si-Al
ternary diagram. These included: (1) an Al-rich intermixed phase for the
binary systems; and (2) an amorphous phase mainly containing Ca, Si and
Al for the ternary systems. After longer curing periods (>180 days), STEM
revealed the formation of an unidentified nano-crystalline phase richer in
Fe with a distinct rod/needle-like morphology and a d-spacing of
approximately 7.2 Å. The evolution of the Mössbauer spectra could
suggest Fe oxidation from the 2+ to 3+ state during slag hydration. This
was accompanied by a shift from tetrahedral to octahedral coordination as
a new Fe-containing hydrated phase is formed. There was no substantial
evidence of Fe uptake in the more abundant hydrated phases, such as in
ettringite or monosulfoaluminates, contrary to the models obtained from
thermodynamic simulations and synthetic systems.
To conclude, the overall properties of the blends containing the industrial
residues support their potential for valorization as SCM in CAC-based
systems. However, it will be critical to continue this investigation to clearly
identify the Fe-bearing hydrated phase/s and to, ultimately, utilize the
understanding of the hydration mechanism in improving
performance.
iv ASTOVEZA 2022
________________________________________________ Samenvatting
Samenvatting
De druk om de cementindustrie koolstofarm te maken in het licht van
duurzaamheidsdoelstellingen heeft de afgelopen decennia de zoektocht naar
nieuwesupplementaire cementerende materialen. (SCM) gemotiveerd.
Industriële residuen, die anders op stortplaatsen terechtkomen of enkel
toepassingen kennen met een lage toegevoegde waarde, zijn van bijzonder
belang in bijna-nul-afvalprocesschema's.
Dit proefschrift onderzocht het potentieel voor valorisatie van behandelde
residuen als SCM in calciumaluminaatcement (CAC)-calciumsulfaat
hemihydraat (HH) systeem. Deze residuen omvatten non-ferrometallurgie
(NFM) slakken, bodemas van de gemeentelijke afvalverbrandingsoven
(AEC), jarosiet en papierfabricageresiduen. Met behulp van isotherme
calorimetrie en mechanische sterktetest werd eerst een screeningstest
uitgevoerdom hun reactiviteit op jonge en oude leeftijd te vergelijken. De zeer
amorfe NFM-slakken vertoonden een superieure reactiviteit, waaruit blijkt dat
ze op lange termijn (> 28 dagen) bijdragen aan de mechanische sterkte van
mortels bij een vervangingsniveau van 30 gew.%. De langzame
hydratatiekinetiek van de slakken in de eerste dagen van hydratatie kan echter
een probleem zijn voor CAC-systemen, aangezien hun toepassingen sterk
afhankelijk zijn van hun snelle uithardingseigenschappen. Bovendien
onderscheiden de amorfe aard van de NFM-slakken en de dominantie van ijzer
(Fe) in hun samenstelling ze van traditionele SCM's. Op het moment van
schrijven zijn er slechts beperkte publicaties die het hydratatiemechanisme
van Fe in op CAC gebaseerde bindmiddelen bespreken.
Om de langzame kinetiek van de hydratatie van de slak te overwinnen zijn drie
factoren onderzocht, waaronder: (1) het gehalte aan sulfaten door de
CAC/HH-verhouding te variëren; (2) de toevoeging van 5 gew.% kalkrijk
papierresidu als activator; en (3) de opname van 30 gew.% PC in ternaire
CAC-HH-PC-systemen. Voor alle formuleringen werden de vroege
hydratatiereacties gekarakteriseerd met behulp van isotherme calorimetrie op
pastamonsters. De maatvastheid en mechanische sterkte van standaardmortels
werd gemonitord over een uithardingsperiode van 1 jaar. Verder werd in deze
studie een methodologie ontwikkeld om de hydratatiegraad van de slak te
schatten met behulp van een niet-destructieve techniek op basis van X-stralen
computertomografie (XCT) in combinatie met volume-analyse. Verschillende
technieken werden gebruikt om de evolutie van de fase-assemblage te volgen,
waaronder: röntgendiffractie (XRD), thermogravimetrische analyse (TGA),
verzwakte totale reflectie-Fourier-transformatie infraroodspectroscopie
(ATR-FTIR). Ten slotte werden ook transmissie-elektronenmicroscopie
(TEM), scanning transmissie-elektronenmicroscopie (STEM) en Mössbauer-
spectroscopie uitgevoerd om een beter begrip te verwervenover het lot van Fe,
het hoofdbestanddeel van de slak, gedurende de hydratatie.
ASTOVEZA 2022 v
Samenvatting ________________________________________________
Voor alle formuleringen met een vervangingsniveau van 30 gew.% NFM slak
werd een hoge vroege sterkte bekomen. Het verhoogde gehalte aan sulfaten
(CAC/HH-verhouding = 1,6) verbeterde de mechanische sterkte door een
verhoogde ettringiet vorming. De toevoeging van PC en papierresten
verbeterde ook de sterkte op jonge leeftijd. De toevoeging van slakken
verbeterde de maatvastheid van mortels door het bindmiddel te verdunnen,
waardoor de overmatige uitzetting in de ternaire systemen werd verminderd.
De XCT-resultaten toonden aan dat ongeveer 50 gew.% van de slak reageerde
na meer dan 180 dagen.. Dit was in overeenstemming met de schattingen
verkregen met Mössbauer-spectroscopie, gebaseerd op de relatieve
hoeveelheden absorptiegebieden van de nieuw gevormde Fe-soorten tijdens
de hydratatie. Hoewel het gehalte aan sulfaten geen duidelijke invloed had op
de oplossnelheid van de slak op basis van de XCT-resultaten, leek de
toevoeging van PC en papierresten dit in de eerste 28 dagen te hebben
versneld. Op langere termijn bleef de reactiegraad echter relatief vergelijkbaar
met die verkregen voor de binaire CAC-HH-systemen. Met behulp van de
ATR-FTIR-spectra werd een hypothetische piekverplaatsing van de Si-O
asymmetrische rekband van de slak bij ongeveer 930 cm-1 waargenomen, wat
deze reactiviteit ondersteunt.
Wat het lot van Fe betreft, leverde TEM bewijs aan voor de opname in
gehydrateerde fasen gegroepeerd in twee samenstellingsclusters op een Ca-Si-
Al ternair diagram. Deze omvatten: (1) een Al-rijke vermengde fase voor de
binaire systemen; en (2) een amorfe fase die voornamelijk Ca, Si en Al bevat
voor de ternaire systemen. Na langere uithardingsperioden (>180 dagen),
onthulde STEM de vorming van een niet-geïdentificeerde nanokristallijne fase
rijker aan Fe met een duidelijke staaf/naaldachtige morfologie en een d-
afstand van ongeveer 7,2 Å. De evolutie van de Mössbauer-spectra zou
kunnen wijzen op de oxidatie status van Fe (van 2+ naar 3+) tijdens de
hydratatie van de slak. Dit ging gepaard met een verschuiving van
tetraëdrische naar octaëdrische coördinatie als een nieuwe Fe-bevattende
gehydrateerde fase werdgevormd. Er was geen substantieel bewijs van Fe-
opname in de meer overvloedige gehydrateerde fasen, zoals in ettringiet of
monosulfoaluminaten, in tegenstelling tot de modellen verkregen uit
thermodynamische simulaties en synthetische systemen.
Globaal gezien, de resultaten in dit werk ondersteunen het
valorisatiepotentieel van industiële residuen als SCM in op CAC gebaseerde
systemen. Het zal echter van cruciaal belang zijn om dit onderzoek voort te
zetten om de Fe-dragende gehydrateerde fase(s) duidelijk te identificeren en
om uiteindelijk het begrip van het hydratatiemechanisme te gebruiken bij het
verbeteren van de prestaties van de mengsels.
vi ASTOVEZA 2022
______________________________________________________ Résumé
Résumé
Au vu des objectifs de durabilité visant à décarboniser l'industrie du
ciment, les liants alternatifs ont
l'échelle mondiale au cours des dernières décennies. Ces liants sont basés
-produits industriels comme substituts aux ciments,
ou comme granulats ou encore comme composants de matériaux de
construction, à performances équivalentes.
Dans le cadre de cette thèse, des résidus traités valorisables sont évalués
comme addition cimentaire notamment dans des mélanges binaires
(ciment alumineux CAC - hémihydrate HH). Trois des sous-produits ont
fait l'objet d'une évaluation plus approfondie : i) un laitier riche en fer
provenant d'une usine pilote de plomb-zinc sous forme granulée (trempe)
ii) un résidu de jarosite ; iii) un résidu du recyclage du papier. Pour
identifier et sélectionner ces résidus, il a été réalisé un test de
conductimétrie pour mesurer la cinétique de dissolution; des mesures de
flux de chaleur pour apprécier les réactions au jeune âge; et une mesure de
résistance mécanique sur mortiers standards pour caractériser la réactivité
à
industriels traités valorisables et son impact positif sur les propriétés des
ciments composés à base de CAC. Néanmoins, il reste encore des
questions importantes en suspens, tant pour expliquer les mécanismes
fondamentaux de l'hydratation que pour tester la durabilité de ces liants.
Ainsi, les travaux futurs sur l'ajout de laitier devraient se concentrer sur des
analyses complémentaires pour identifier la phase nanocristalline
contenant du fer.
Introduction
Le projet européen MSCA-ETN SOCRATES a pour objectif de
promouvoir l'économie circulaire en créant des procédés de valorisation
des résidus industriels qui génèrent le moins possible de nouveaux déchets.
Comme le montre la Figure 1, la technologie SOCRATES vise tout
dernière voie de valorisation qui est étudiée dans le cadre de cette thèse.
Les additions prétraitées devant être ajoutées au ciment sont appelées
résidus traités valorisables ou encore « sous-produits SOCRATES ».
Figure 1 Schéma simplifié du projet SOCRATES axé sur les travaux de l'auteure:
valorisation des résidus comme additions cimentaires dans les ciments composés.
ASTOVEZA 2022 ix
Thesis Summary in French _____________________________________
dans le monde;
-50 % de la demande mondiale de
béton est prévu
Au vu de ces chiffres, on comprend pourquoi les liants alternatifs ont fait
x ASTOVEZA 2022
____________________________________ Thesis Summary in French
CAC que pour les liants riches en ciment Portland, tout en maintenant le
même niveau de performances. À ce jour, il existe peu d'articles
les liants à base de CAC. Les travaux de recherche publiés se limitent aux
additions traditionnelles, principalement les laitiers de hauts-fourneaux et
les cendres volantes [13 16].
ASTOVEZA 2022 xi
Thesis Summary in French _____________________________________
Objectifs et Problématique
Deux objectifs principaux ont été définis dans le cadre du projet
SOCRATES :
1. Évaluer la réactivité des sous-produits SOCRATES dans les
systèmes à base de CAC ;
2. Proposer des techniques d'activation de ces sous-produits afin
d'améliorer la cinétique de leur réaction, et les performances
associées.
Guidée par ces objectifs, une méthodologie a été mise en place afin de
repérer le résidu ayant le potentiel de valorisation le plus élevé. L'évolution
des propriétés physiques et chimiques du résidu sélectionné au cours de
l'hydratation a ensuite été étudiée en détail.
Matériaux et Méthodologie
Six sous-produits SOCRATES ont été étudiés : un laitier riche en fer
provenant d'une usine pilote de plomb-zinc, disponible sous forme
d'entre eux ont fait l'objet d'une évaluation plus approfondie. Ces trois
résidus sont représentés Figure 3.
Figure 3 Les principaux sous-produits SOCRATES évalués dans cette étude, notamment :
le laitier riche en fer granulé (a) ; le résidu jarosite (b) ; et les résidus du recyclage du
papier (c)
ition (le
laitier) est remplacée par un volume égal de quartz broyé. Les pâtes de
ciment ont été préparées avec un rapport eau/liant (e/l) de 0,65 et stockées
sous enveloppes hermétiques à 20°C. Les mortiers standards ont un rapport
e/l de 0,5 et ils ont été conservés sous eau à 20°C.
Plusieurs techniques ont été utilisées pour caractériser les propriétés des
mélanges (entre 1 jour et 1 an), et certaines ont été spécifiquement adaptées
aux propriétés particulières du laitier et à sa teneur élevée en fer (i.e.
densité élevée, masse atomique élevée, effet Mössbauer important) :
la résistance mécanique et la stabilité dimensionnelle ont été
suivies en utilisant des éprouvettes de mortier;
le degré d'hydratation du laitier a été quantifié à l'aide de la
microscopie électronique à balayage (MEB), de la tomographie
par rayons X (XCT) et de la spectroscopie Mössbauer;
l'assemblage des phases en fonction de la proportion de CAC a été
simulé à l'aide de la modélisation thermodynamique et plusieurs
techniques, notamment la calorimétrie isotherme, la diffraction des
rayons X (DRX), l'analyse thermogravimétrique (ATG) et la
spectroscopie infrarouge à transformée de Fourier (IRTF) ont été
utilisées pour suivre expérimentalement l'évolution des
assemblages;
le mécanisme d'hydratation du fer contenu dans le laitier a été
décrit à partir des résultats de la spectroscopie Mössbauer, de la
microscopie électronique à transmission (MET) et de la
microscopie électronique à transmission à balayage (METB).
Résultats et Discussion
Pour un taux de remplacement massique de 30 % de laitier issu de métaux
non-ferreux, l'indice d'activité (IA) déduit de la résistance mécanique sur
Propriétés physiques
Les mortiers incorporant le laitier issu de métaux non-ferreux granulé et
1,3 % d'acide citrique (Figure 4) ont atteint des valeurs de résistance à la
ASTOVEZA 2022 xv
Thesis Summary in French _____________________________________
Figure 5 Photos des mortiers B et B-PC après 45 jours de conservation sous eau
ra
phase
hydratée contenant du fer.
Figure 8 Images MET de certaines phases observées dans les pâtes contenant du laitier,
notamment : gels contenant du fer (a-b) avec leur diffractogramme (c) ; ettringite (d) ;
strätlingite (e) et wüstite (f)
À plus longs termes, les résultats obtenus par microscopie électronique ont
montré que la composition des gels tend à se déplacer, dans le diagramme
révéler avec certitude l'identité de cette phase hydratée riche en fer. Des
techniques ou des méthodologies complémentaires sont nécessaires pour
xx ASTOVEZA 2022
____________________________________ Thesis Summary in French
Conclusion et perspectives
La caractérisation approfondie des propriétés physiques et chimiques des
amorphe riche en fer dans des liants binaires CAC-HH et ternaires CAC-
HH-PC. La méthodologie développée (tomographie, etc.) a permis de
mon -delà de 180 jours. Ce
résultat a été confirmé par spectroscopie Mössbauer, qui a montré
également que l'hydratation du laitier s'accompagnait de l'oxydation du fer
2+ à 3+, et de la formation de nouveaux hydrates contenant du fer. À l'aide
de la microscopie électronique (MET et METB), le fer a été détecté au
jeune âge dans des phases complexes formées de gels amorphes riches en
aluminium. Le fer a également été détecté à des proportions plus élevées à
plus long terme, au sein une phase nanocristalline non identifiée. Dans les
systèmes ternaires, cette phase était également associée à du calcium, du
silicium et une petite quantité de magnésium.
L'effet du ciment portland, des sulfates et des résidus du recyclage du
papier sur la réactivité du laitier ont également été explorés dans cette
étude. Les résultats ont montré qu'il n'y avait pas de différences
substantielles dans le degré d'hydratation du laitier entre les systèmes à
faible et à haute teneur en sulfate étudiés. Au contraire, l'ajout de ciment
portland et de résidus du recyclage du papier ont montré une augmentation
du degré d'hydratation, toutefois cantonnée au jeune âge (< 28 jours). Cet
2+
dans le système,
ainsi que de l'augmentation de l'alcalinité favorisant la dissolution du
laitier.
Pour conclure, il reste encore des questions importantes en suspens, tant
pour expliquer les mécanismes fondamentaux de l'hydratation que pour
tester la durabilité de ces liants. Ainsi, les travaux futurs devraient se
concentrer sur des analyses complémentaires pour identifier la phase
nanocristalline contenant du fer. Elles pourraient utiliser des techniques de
caractérisation supplémentaires, ou bien reproduire les systèmes de
manière synthétique (phases pures) pour limiter notamment le problème
Bibliographie
[1] United Nations Environement Program (UNEP), Buildings and
climate change: Summary for decision-makers, 2009,
www.unep.fr/scp/sun.
[2] K.L. Scrivener, V.M. John, E.M. Gartner, Eco-efficient cements:
Potential economically viable solutions for a low-CO2 cement-based
materials industry, Cement and Concrete Research 114 (2018) 2 26.
[3] P.J.M. Monteiro, S.A. Miller, A. Horvath, Towards sustainable
concrete, Nature materials 16 (2017) 698 699.
[4] A. Vollpracht, B. Lothenbach, R. Snellings, J. Haufe, The pore
solution of blended cements: a review, Mater Struct 49 (2016) 3341
3367.
[5] S. Samad, A. Shah, M.C. Limbachiya, Strength development
characteristics of concrete produced with blended cement using ground
granulated blast furnace slag (GGBS) under various curing conditions,
1213.
[6] SOCRATES EU MSCA-ETN, Project proposal: Innovative training
networks (ITN) call: H2020-MSCA-ITN-2016, Part B, 2016, https://etn-
socrates.eu/.
[7] J.E. Rossen, Composition and morphology of C-A-S-H in pastes of
alite and cement blended with supplementary cementitious materials,
Lausanne, EPFL, 2014.
[8] R. Snellings, G. Mertens, J. Elsen, Supplementary Cementitious
Materials, Reviews in Mineralogy and Geochemistry 74 (2012) 211 278.
Table of Contents
Acknowledgements .............................................................................. i
Abstract .............................................................................................. iii
Samenvatting ....................................................................................... v
Résumé .............................................................................................. vii
List of Figures ................................................................................ xxxi
List of Tables ................................................................................ xxxix
Glossary............................................................................................. xli
Introduction ....................................................................................... 1
Chapter 1: State-of-the-art ............................................................... 9
1.1. The SOCRATES Project............................................................ 13
1.2. Calcium aluminate blended cements ......................................... 17
1.2.1. Pure CAC systems .............................................................. 18
1.2.2. Binary CAC-C$ systems ..................................................... 21
1.2.3. Ternary CAC-PC-C$ systems ............................................. 24
1.3. Supplementary Cementitious Materials ..................................... 27
1.3.1. Traditional SCM .................................................................. 27
1.3.2. Non-conventional SCM ....................................................... 30
1.4. The role of Fe in inorganic polymers and cement binders ......... 33
1.4.1. Fe speciation in slag-containing inorganic polymers ........ 33
1.4.2. Fe speciation in cementitious binders ................................ 34
1.4.3. Other techniques to characterize and quantify the reactivity
of slags ........................................................................................... 37
Chapter 2: Materials, methods and residue selection .................. 43
2.1. Materials and formulation .......................................................... 45
2.1.1. The Socrates residues ......................................................... 45
2.1.2. Cements .............................................................................. 48
2.1.3. Other blend components ..................................................... 48
2.1.4. Dry mix formulations .......................................................... 48
2.2. Methods ..................................................................................... 50
2.2.1. Basic characterization of raw materials ............................ 50
xxvi ASTOVEZA 2022
Table of Contents ______________________________________________
List of Figures
Figure 1 Simplified flow sheet of the SOCRATES Project .................... 1
Figure 2 Chemical compatibility of some of the SOCRATES residues .. 3
Figure 3 Some applications of CAC-based binders: ............................... 5
Figure 2.30 SEM-SE images of (a) no jarosite and (b) 100JAR paste
sample..................................................................................................... 75
Figure 2.31 Comparison of the ion concentrations after the dissolution test
................................................................................................................ 76
Figure 3.1 Initial and final setting time of formulation A and its references
................................................................................................................ 84
Figure 3.2 Compressive strength of formulation A (primary axis) and its
SAI evolution (secondary axis) .............................................................. 85
Figure 3.3 Flexural strength of formulation A (primary axis) and its SAI
evolution (secondary axis)...................................................................... 85
Figure 3.4 Length variation of formulation A and its reference ............ 86
Figure 3.5 Compressive strength of formulation B (primary axis) and its
SAI evolution (secondary axis) .............................................................. 88
Figure 3.6 Flexural strength of formulation B (primary axis) and its SAI
evolution (secondary axis)...................................................................... 88
Figure 3.7 Length variation of formulation B and its references .......... 88
Figure 3.8 (a) SEM-BSE micrograph of formulation A mortar after 90
days ........................................................................................................ 90
Figure 3.9 (a) SEM-BSE micrograph of formulation A reference cement
................................................................................................................ 91
Figure 3.10 SEM-BSE micrograph of formulation A pastes ................. 91
Figure 3.11 Illustration of the steps involved in the volume analysis of
XCT ........................................................................................................ 93
Figure 3.12 Estimated slag volume and hydration degree using XCT .. 94
Figure 3.13 Illustration of XCT analysis using cylindrical volumes ..... 95
Figure 3.14 Estimated (a) slag total volume and (b) slag hydration degree
................................................................................................................ 96
Figure 3.15 Phase assemblage of the reference (no slag) binary binder
calculated from GEMS thermodynamic modelling software ................. 98
Figure 3.16 Phase assemblage of the binary binder incorporating 30% slag
with varying CAC and HH proportions ................................................. 98
Figure 3.17 Heat flow measured ex-situ using isothermal calorimetry . 99
Figure 3.18 Cumulative heat release measured using isothermal
calorimetry ........................................................................................... 100
Figure 3.19 XRD patterns of formulation A (solid black lines) .......... 101
Figure 3.20 XRD patterns of formulation B (solid blue lines) ............ 103
Figure 3.21 Evolution of the DTG curves of (a) formulation A and (b)
formulation B........................................................................................ 104
Figure 3.22 Total bound water content of formulations A and B ....... 104
Figure 3.23 ATR-FTIR spectra of formulation A (solid black lines). . 105
Figure 3.24 ATR-FTIR spectra of formulation B (solid blue lines). ... 106
Figure 3.25 TEM images and the atomic composition of areas in the (A)
.............................................................................................................. 108
Figure 3.26 TEM images and the atomic composition of areas in the (A)
.............................................................................................................. 109
Figure 3.27 TEM images, diffraction patterns and the atomic composition
of Fe-containing gels in the (A) 4.5CAC-HH paste sample ................. 110
Figure 3.28 TEM images before and after EDS analysis of Fe-containing
gels in the (A) 4.5CAC-HH paste sample ............................................ 110
Figure 3.29 TEM images and the atomic composition of crystalline phases
in the (A) 4.5CAC-HH paste sample .................................................... 110
Figure 3.30 TEM images and the atomic composition of marked areas in
the (B) 1.6CAC-HH paste sample ........................................................ 111
Figure 3.31 STEM elemental mapping, diffraction patterns ............... 113
Figure 3.32 STEM elemental mapping, diffraction patterns ............... 114
Figure 3.33 Ternary diagram of Fe-containing hydrate gels with 1-12% Fe
.............................................................................................................. 115
Figure 3.34 STEM image focused on an Fe-rich region...................... 116
Figure 3.35 STEM image focused on an Fe-rich region (a); along with its
corresponding annular dark field image (b); Fe map from EDS analysis (c);
and the diffraction pattern .................................................................... 116
Figure 3.36 STEM images focused on the needle-like Fe-rich phase . 117
Figure 3.37 TEM images of needle-like Fe hydroxide phases ............ 118
Figure 3.38 TEM images of (a) Al(OH)3 with its inset selected-area
electron diffraction ............................................................................... 119
Figure 3.39 57Fe Mössbauer spectrum evolution of formulation A ..... 121
Figure 3.40 57Fe Mössbauer spectrum of formulation A and its
components ........................................................................................... 123
Figure 3.41 Absorption areas (AA) calculated from the two modelling
approaches ............................................................................................ 127
Figure 3.42 Slag hydration degree in formulation A paste.................. 128
Figure 3.43 (a) Heat flow and (b) cumulative heat release of pastes
containing 23 wt% paper residue (in blue) and 30 wt% slag (in black).129
Figure 3.44 Initial and final setting time of formulations A-TC; reference
no TC; and formulation A .................................................................... 130
Figure 3.45 Compressive strength (primary axis) and SAI evolution
(secondary axis) of formulation A-TC (blue), its reference formulation
(gray) and formulation A (black). ........................................................ 131
Figure 3.46 Flexural strength (primary axis) and SAI evolution (secondary
axis) of formulation A-TC (blue), its reference formulation (gray) and
formulation A (black). .......................................................................... 131
Figure 3.47 Length variation of formulations A-TC (blue) and A (black)
.............................................................................................................. 132
Figure 3.48 The slag hydration degree (primary axis) and slag total volume
of formulations A-TC and A estimated from XCT + volume analysis 132
Figure 3.49 (a) XRD patterns; (b) Mössbauer spectra; (c) DTG curves; and
(d) total bound water evolution of formulations A-TC ........................ 133
Figure 3.50 TEM images and the atomic composition of marked areas in
the (A-TC) 4.5CAC-HH-5TC paste sample ......................................... 134
Figure 4.1 Initial and final setting time of formulation A-PC, B-PC .. 140
Figure 4.2 Compressive strength of formulation A-PC (primary axis)141
Figure 4.3 Flexural strength of formulation A-PC (primary axis)....... 141
Figure 4.4 Length variation of formulation A-PC and its references .. 142
Figure 4.5 Compressive strength of formulation B-PC (primary axis) 144
Figure 4.6 Flexural strength of formulation B-PC (primary axis) ....... 144
Figure 4.7 Length variation of formulation B-PC and its references .. 146
Figure 4.8 Pictures of A-PC and B-PC mortars with their references . 147
Figure 4.9 Illustration of the steps involved in the volume analysis of XCT
.............................................................................................................. 148
Figure 4.10 Estimated slag volume and hydration degree using XCT 149
Figure 4.11 Phase assemblage of the reference ternary CAC-HH-PC
binder.................................................................................................... 152
List of Tables
Table 3.1 Estimated slag area and hydration degree using SEM-BSE .. 92
Table 3.2 FTIR band assignments based on the distinct spectral peaks107
Table 3.3 57Fe Mössbauer hyperfine parameters used for Model 1. .... 124
Table 3.4 57Fe Mössbauer hyperfine parameters used for Model 2. .... 124
Table 4.1 FTIR band assignments based on the distinct spectral peaks161
Table 5.1 Composition of binders incorporating the Fe-rich slag ....... 181
Table 5.2 Consolidated phase assemblage based on thermodynamic
modelling.............................................................................................. 184
Glossary
Cement shorthand notations for oxides:
H: H2O C: CaO A: Al2O3 S: SiO2 F: Fe2O3 $: SO3 c: CO2 T: TiO2
Materials:
Calcium aluminate cement, Ciment Fondu® or Secar
CAC 51®
PC Ordinary Portand cement, Milke® Premium
HH -hemihydrate, Prestia Selecta®
C$ Calcium sulfate
(quartz) filler Fine sand, Millisil E400®
NFM slags Non-ferrous metallurgy slags
slag Granulated iron-rich slag (Chapters 3and 4)
paper residue Residue from paper recycling, Top-Crete®
SCM Supplementary cementitious materials
binder Cement (CAC, PC) + HH + SCM or filler
IP Inorganic polymer
Techniques:
PSD Particle size distribution
SSA Specific surface area
XRF X-ray fluorescence
XRD X-ray diffraction
TGA Thermogravimetric analysis
ATR- Attenuated total reflection-
FTIR Fourier transform infrared spectroscopy
XCT X-ray computed tomography
SEM Scanning electron microscopy
(Scanning) Transmission electron
(S)TEM microscopy
MS Mössbauer spectroscopy
XANES X-ray absorption near edge structure
NMR Nuclear magnetic resonace spectroscopy
XAS X-ray absorption spectroscopy
Other abbreviations:
wt% weight percent
at% atomic percent
vol% volume percent
w/b water to binder ratio
SAI strength activity index
VOI volume of interest in XCT
BSE backscaterred electron in SEM or TEM
EDS energy-dispersive (detector) in SEM or TEM
primary axis located on the vertical left
secondary axis located on the vertical right
Introduction
Circular economy is one of the fundamental building blocks of the
2021 European Green Deal [1]. This concept revolves around regenerative
resource cycles with a focus on minimization of waste through process
optimization, materials substitution and recycling. As a consequence,
research and development of sustainable processes and products is
1
-Curie Actions (MSCA) European Training Network (ETN) for the
sustainable, zero-waste valorization of critical-metal-containing industrial process residues under the
EU Framework Program for Research and Innovation Horizon 2020 Grant Agreement No.721385
ASTOVEZA 2022 1
Introduction _________________________________________________
-
Generally, when by-products are adapted as substitutes to cements, as
aggregates, or as precursors for other construction materials,
environmental and economic benefits are envisioned while maintaining
acceptable product performance [3,4]. If the SOCRATES residues prove
to be viable candidates, the impact of this valorization path would be
substantial, considering the following figures [5]:
8 % of the global anthropogenic CO2 is attributed to cement
manufacturing
30 billion metric tons of concrete is produced globally per year
25-50 % of projected global increase in demand for concrete in
2050
With these, it is easy to comprehend how sustainable binders have received
significant attention in Europe in the recent decades. Nonetheless, there
remain plenty of research gaps to be addressed not only in terms of the
variety of residues and cement systems investigated, but also in
understanding the properties and hydration mechanism of these binders.
Majority of the previous research works are focused on the addition of the
more traditional SCM (i.e., GGBFS2, coal combustion fly ash, and
limestone) in ordinary Portland cement (PC)-dominated binders [6,7].
hydration by acting as an inert filler in the early hours. This was due to the
enhanced nucleation and growth when the SCM added is finer than the
cement component; and to the higher effective water to cement ratio as less
proportion of cement is added in the blend [8,9]. In the early stage of
hydration, this effect is said to be mainly physical in nature considering the
often lower dissolution rate of SCM relative to the cement component. At
long term however, ions released from the dissolving SCM could
chemically alter the hydration mechanism. For instance, siliceous SCM
such as fly ashes contribute silica for the formation of strätlingite, C2ASH8.
In addition, the dissolution of lime and alkali from the SCM could also
2
ground granulated blast furnace slag (GGBFS)
2 ASTOVEZA 2022
_________________________________________________ Introduction
modify the pH of the pore solution, thereby modifying the solubility of the
phases during hydration [6,9].
ASTOVEZA 2022 3
Introduction _________________________________________________
For instance, whereas the latter consist mainly of CaO, SiO2 and Al2O3,
NFM slags and jarosite residue are much richer in iron and sulfates,
respectively. The difference in the chemistry among the SOCRATES
residues, traditional SCM and cements is depicted in a ternary diagram in
Figure 3. Although this diagram is already valuable in grouping the
different resources, there is more information not disclosed. For instance,
the phase where iron partitions in the NFM slag and jarosite is amorphous
versus crystalline, and predominantly Fe2+ versus Fe3+.
Although MSWI bottom ashes could have major (>70 wt%) proportion of
CaO, SiO2 and Al2O3, they typically have a much lower degree of
vitrification as opposed to fly ashes or the GGBFS. Apart from the
difference in chemistry and oxidation state of certain metals (that might
affect the glass forming ability), this is also due to the lower firing
temperature they have been exposed to (where only partial vitrification
occurred) as well as due to the slower cooling rate afterwards. On the other
hand, the substitution level with the commercially-marketed paper residue
[14] remains limited due to its relatively high free lime content.
Overall, the NFM slags are of particular interest as SCM among the
SOCRATES residues. This is mainly because of their expected highly
amorphous nature which is often linked to their reactivity [15]. In addition,
these slags have undergone high temperature pre-treatment such that the
heavy metals and toxic components are substantially depleted. However,
4 ASTOVEZA 2022
_________________________________________________ Introduction
3
The ability of the binder to fully consume the mix water for hydration
ASTOVEZA 2022 5
Introduction _________________________________________________
compare the reactivity of the residues. It ends with a short case study
evaluating the use of jarosite as a possible sulfate source in the binary
CAC-HH blend. This chapter aims to address the first objective on the list
by demonstrating the superior reactivity of one of the Fe-rich NFM slags
among the residues tested. This selection narrowed down the focus of the
last two chapters of the manuscript.
The reactivity of the Fe-rich slag in the binary (CAC-HH) and ternary
(CAC-HH-PC) systems are presented in Chapters 3 and 4, respectively. A
rapid non-destructive technique for quantifying the degree of slag
hydration using tomography is demonstrated in these chapters. The
physical properties (strength, setting time, and dimensional stability) of the
binders are evaluated using standard mortar samples. The phase
assemblage evolution is followed from the early hours (isothermal
calorimetry) up to 1 year of curing (XRD, TGA, FTIR) using paste
samples. The results were compared to models obtained from (GEMS)
thermodynamic simulation. Finally, the fate of Fe during hydration is
described at the end of both chapters using advanced microscopy and
spectroscopy techniques. The second objective of the research is addressed
by examining the potential role as activators of the PC (in the ternary
system), sulfates (by varying the CAC/HH ratio), and lime (in the form of
the paper residue).
ASTOVEZA 2022 7
____________________________________________________ Chapter 1
Chapter 1
State-of-the-art
Contents
Introduction .. 11
Conclusion ... 41
ASTOVEZA 2022 9
____________________________________________________ Chapter 1
Introduction
This chapter provides a wider outlook on the SOCRATES Project; calcium
aluminate blended cements; supplementary cementitious materials (SCM)
in calcium aluminate cement (CAC)-based systems; and the reactivity of
non-ferrous metallurgy (NFM) slags. It is essential to begin the discussion
with the origin of the SOCRATES residues in order to understand their
properties. This also helps assess the environmental impact of their
valorization as SCM in blended cements. This discussion is followed by
the state-of-the-art on the properties and hydration mechanism of CAC-
based blends in pure CAC, binary CAC-calcium sulfate (C$) and ternary
CAC-C$-Portland cement (PC) systems.
Moreover, an overview of the incorporation of traditional SCM and
industrial by-products (bottom ash, jarosite and NFM slags) in cements is
provided highlighting the need for the latter alternative materials and the
state-of-the-art in this application. The discussion in this particular section
n Appendix A. Finally,
a brief literature review on the reactivity of Fe-rich NFM slags in cement
and inorganic polymer (IP) systems is provided at the end of the chapter.
Among the SOCRATES residues, the NFM slags are of particular interest
considering their relatively low heavy metals content and their amorphous
nature. Both of these are generally considered as favorable properties of
SCM and will be elaborated in more detail in this section. The discussion
on the fate of Fe during hydration will provide a critical background for
the interpretation of results in Chapters 3 and 4 which are dedicated to the
addition of NFM slags in CAC-based binders.
ASTOVEZA 2022 11
____________________________________________________ Chapter 1
It has been reported that there are between 150,000 and 500,000 landfills
throughout EU which could be a potential source of secondary raw
materials and energy [27]. The benefits from their extraction do not only
range from economic (minimizing the landfill remediation cost) and
ecologic (land reclamation and health hazards from associated pollution)
aspects but could also trigger novel technologies to facilitate the metal
extraction [28]. Case in point, a wide range of metallurgical processes
including plasma-, bio-, solvo-, electro- and ionometallurgy are explored
in the SOCRATES project both in extracting valuable metals and in
ASTOVEZA 2022 13
Chapter 1 ____________________________________________________
removing the hazardous components of the residues [29 33]. Some of the
industrial by-products investigated include: NFM slags, MSWI bottom
ashes, jarosite and paper-fabrication residues. Most of which have been
left in stockpiles or landfills for a long period of time for the lack of suitable
application and some due to their toxic components.
The NFM slags in this study are coming from secondary copper and lead-
zinc smelting processes typically operated at 1200-1300°C in reducing
environment. They are generally dominated by FeO and SiO2, and contain
moderate amounts of Al2O3 and CaO. The degree of vitrification is
determined by the cooling rate following the tapping4. Granulation is often
performed by quenching in water to yield higher glass fraction as opposed
to air-cooling conditions. When slowly-cooled, the mineralogy is often
dominated by fayalite (Fe2SiO4) along with other olivine and spinel group
phases. Compared to ferrous metallurgy slags, NFM slags are typically
richer in trace elements (i.e., Cu, Cr, Mn, Pb and Zn) that may pose
challenge to their reuse in construction applications. Case in point, heavy
metals have been reported to retard the precipitation of portlandite with the
reduction of pH resulting from the hydrolysis of metal ions [34].
One of the promising technologies to refine NFM slags, particularly
fayalitic copper slags, in the project is the submerged plasma fuming
process (Figure 1.2). It aims to recover Pb and Zn, and remove other
volatile components (i.e., Cl, F, Ge) using an energy-dense reducing gas
made from a combination of heated compressed air and natural gas. The
volatile components are precipitated in a post combustion duct and
recovered in the form of oxide dusts [35]. The resulting slag, being
generally metal-depleted, is being marketed since 2018 as a clean synthetic
mineral for sustainable building applications [36].
Another clean (metal-depleted) commercialized by-product assessed in
this project is the paper residue. It is derived from the de-inking sludge
during the recycling of paper wastes in a fluidized-bed incinerator. This
process is designed for waste-to-energy applications, such that
approximately 7 GWh of electricity (or equivalent to the estimated annual
energy requirement of 151,000 households) is generated annually from the
incineration of 140,000 tons of paper pulp residue in this process alone
[37]. The remaining lime-rich residue after the combustion is marketed as
a green binder displaying pozzolanic activity that can be applied in dosages
up to 20 wt% [14].
4
removal of slag by allowing it to flow in a tap hole at the lower region of a blast furnace
14 ASTOVEZA 2022
____________________________________________________ Chapter 1
ASTOVEZA 2022 15
Chapter 1 ____________________________________________________
5
in particular: selective catalytic reduction of NOx with NH3 ; propane dehydrogenation to
propylene and propylene epoxidation
6
kaolinite, fly ash and blast furnace slags more conventionally used as precursors to IP
16 ASTOVEZA 2022
____________________________________________________ Chapter 1
ASTOVEZA 2022 17
Chapter 1 ____________________________________________________
refractory and decorative usage, while the former is more adapted for
higher volume construction purposes [18,19].
CAC is generally dominated7 by monocalcium aluminate (CA) and could
alternatively contain significant proportions of calcium di-aluminate (CA2)
and mayenite (C12A7
gehlenite (C2AS) and ferrite phase (C4AF) are typically present and the rest
is principally balanced by smaller percentages (<5 wt%) of belite (C2S),
perovskite (CT), and spinel (MgAl2O4).
7
approximately <60 wt% CA for fused CAC and 60-70 wt% CA for sintered CAC
18 ASTOVEZA 2022
____________________________________________________ Chapter 1
aluminate (C3A), ferrite (C4AF) phase, and calcium carbonate (Cc)8, listed
in Equation 1.4 to 1.10 [19,54].
Equation 1.4
Equation 1.5
Equation 1.6
Equation 1.7
Equation 1.8
Equation 1.9
Equation 1.10
8
originating either as a component of the raw cement; or produced as a carbonation product of C
and CH with the CO2 gas from the atmosphere during curing/sample preparation
ASTOVEZA 2022 19
Chapter 1 ____________________________________________________
Equation 1.11
Equation 1.12
Figure 1.4 SEM images of the hexagonal CAH10 (a) and cubic C3AH6 (b) [65]
20 ASTOVEZA 2022
____________________________________________________ Chapter 1
Equation 1.14
Equation 1.15
Equation 1.16
Equation 1.17
Equation 1.18
Equation 1.19
In high sulfate systems (CAC/C$ < 2), the ettringite formed in the early
hours of hydration remains stable overtime. However, the excessive
ASTOVEZA 2022 21
Chapter 1 ____________________________________________________
Figure 1.5 Heat flow curves of formulations with different proportion of CA and HH
measured per 10 g of paste samples with 0.4 water to binder (w/b) ratio at 20°C using
ex-situ isothermal calorimeter [9]
9
the amorphous nature of AH3 in this formulation hindered its detection using XRD
22 ASTOVEZA 2022
____________________________________________________ Chapter 1
10
ettringite has the lowest (true) density at 1.78 g/cm3 among the other common hydrate phases in
these systems such as monosulfoaluminates (2.01 g/cm3); gibbsite (2.42 g/cm3) and strätlingite (1.94
g/cm3) [78]
ASTOVEZA 2022 23
Chapter 1 ____________________________________________________
crucial role in controlling the setting time of the dry mix. The typical
sulfate carriers (gypsum, anhydrite and HH) readily hydrate and set into
calcium sulfate dihydrate in contact with water leading to an accelerated
rate of expansion and hardening [79]. For this reason, set retarders (i.e.,
citric acid, tartaric acid) are often introduced as admixtures in order to
counteract the influence of sulfates. Nguyen et al.,2019 [82] reported that
at a significant dosage of 2 wt%, citric acid does not only retard hydration
but could also contribute to the conversion of ettringite to
monosulfoaluminate; and influence the mechanical strength of mortars.
This was associated to the formation of citrate ligand around ettringite
crystals that temporarily inhibits its precipitation, consequently favoring
monosulfoaluminate formation [83]. Nonetheless, citric acid is more
typically used in smaller amounts (0.1-0.5 wt%) only to control the setting.
No evident influence to the phase assemblage or to the mechanical strength
should be expected at this dosage [82].
24 ASTOVEZA 2022
____________________________________________________ Chapter 1
Equation 1.20
Figure 1.8 Phase assemblage evolution of: (a) a PC-rich ternary system; and (b) a CAC-
rich ternary system, simulated using GEMS thermodynamic modelling at 20°C
and 1 bar pressure. Phases were abbreviated as follows: (Ms) monosulfoaluminate; (Hc)
hemicarboaluminate; (Mc) monocarboaluminate [19]
In both the (80 wt%) PC- and (65 wt%) CAC-rich formulations modelled
in Figure 1.8, ettringite was predicted to form in a significant quantity and
remained stable throughout the simulated 90 days of hydration. C-S-H and
CH, the main hydration products in pure PC systems, were only present in
the PC-rich system. Instead, the silicates from the minor PC added were
predicted to be incorporated in strätlingite for the CAC-rich formulation.
Furthermore, these models illustrate that CH and AH3 do not co-exist
ASTOVEZA 2022 25
Chapter 1 ____________________________________________________
11
at the amount where ettringite formation is favored over that of Ms
26 ASTOVEZA 2022
____________________________________________________ Chapter 1
of the pore solution, but also by contributing ions (Ca2+, Al(OH)4- and SO42-
) that can participate during hydrate formation. Thermodynamic modelling
estimates pH values above 12 in pore solution of pure PC hydration [90];
whereas a pH less than 12 is expected for CAC-rich systems even reaching
10.5 with excess C$ [9]. This advantage is further supported by the
improved stability of ettringite formed at pH 12 reported in Van-Rompaey,
2006 [91].
ASTOVEZA 2022 27
Chapter 1 ____________________________________________________
12
equivalent to the ASTM C 989 for GGBFS
28 ASTOVEZA 2022
____________________________________________________ Chapter 1
Figure 1.9 Heat flow curves of a pure PC system and blends with 40 wt% replacement
with SCM, calculated per gram of cement [95].
13
pertains to the ability of the SCM to fill up the voids of the binder, associated to its PSD and the
morphology of the particles
14
deleterious expansion in concrete attributed to the formation of hygroscopic sodium silicate gels
from the reaction of highly alkaline cement paste, siliceous aggregates and water
ASTOVEZA 2022 29
Chapter 1 ____________________________________________________
Figure 1.10 (a) Particle size distribution of several SCM compared to that of a PC;
and (b) the compressive strengths in function of cement replacement level, relative to a
pure PC binder [96].
30 ASTOVEZA 2022
____________________________________________________ Chapter 1
ASTOVEZA 2022 31
Chapter 1 ____________________________________________________
15
measure of work needed to perform a process or produce a material
32 ASTOVEZA 2022
____________________________________________________ Chapter 1
the slag. These sustainability measures further support the interest in the
valorization potential, particularly, of NFM slags as SCM. In this regard,
during hydration that could influence the physical and chemical properties
of these binders.
16
The study only took the Fe speciation of the non-amorphous fraction in consideration
ASTOVEZA 2022 33
Chapter 1 ____________________________________________________
significantly within the first 24 h. This timeline was consistent with the
onset of the polymerization exotherm in the calorimetry performed in the
same study [116]. Peys et al., 2019 [115] considered O2 + 4 e- => 2 O2- and
Fe2+ => Fe3++ e- as possible half reactions governing this oxidation, the
result of their study on polymerization under N2 atmosphere revealed
oxygen imbalance in the system. Nonetheless, the study demonstrated that
the observed oxidation was only driven by the atmospheric conditions
during the later stages (>28 days), and not during the first (3) days after
mixing [115].
Figure 1.11 The structure of a synthetic NFM slag proposed in Siakati et al., 2020
[119] simulated based on MD simulations and PDF studies. The slag has a molar
composition of 0.33CaO:0.67FeO-SiO2 following (CaO+FeO)/SiO2 = 1. The [FeOx]
(where x = 5 or 6) and [SiO4] tetrahedra were represented in yellow and dark blue
planes, respectively
34 ASTOVEZA 2022
____________________________________________________ Chapter 1
17
except for the supersulphated slag cement (84.5 wt% slag, 15 wt% anhydrite and 0.5 wt% KOH)
due to the lower amount of Ca versus CEM II and III samples investigated in their study
18
quantum-mechanical simulation based on density functional theory
ASTOVEZA 2022 35
Chapter 1 ____________________________________________________
36 ASTOVEZA 2022
____________________________________________________ Chapter 1
ASTOVEZA 2022 37
Chapter 1 ____________________________________________________
Figure 1.13 ATR-FTIR spectra of three synthetic Fe-rich (65-68 wt% FeO) slags
before (dotted lines) and after (solid lines) 28 days of polymerization modified from
Siakati et al., 2020 [117].
Selective dissolution
In a recent publication, Hallet et al., 2022 [139] adapted the ethylene
diamine tetra acetic acid (EDTA)-triethanolamine (TEA)-NaOH selective
dissolution test originally intended for fly ashes [140], to quantify the
degree of NFM slag reaction in PC-based blended cements. They
determined that almost 25 % of the slag has completely dissolved after 150
days of hydration and that the kinetics were comparable regardless of the
slag-PC substitution level (30 wt%, 50 wt% or 70 wt%) [139]. On the other
hand, a procedure using methanol-salicylic acid has been proposed for
GGBFS-CAC systems in Nakagawa et al., 1990 [141]. However, there has
been no methodology previously published for Fe-rich NFM slag in CAC-
38 ASTOVEZA 2022
____________________________________________________ Chapter 1
Figure 1.14 Figures collated from the quantification of GGBFS hydration degree
using XCT in Wu et al., 2020 [144]. The reconstructed 2D image after the scan is shown
in (a). The grayscale threshold segmentation of the pixels under the green and red arrows
in (a) is illustrated in (b). The 2D (c) and 3D (d) renders after segmentation are also
presented.
ASTOVEZA 2022 39
Chapter 1 ____________________________________________________
40 ASTOVEZA 2022
____________________________________________________ Chapter 1
Conclusion
The SOCRATES project offers potential types of novel SCM in the form
of pre-treated industrial by-products, including: NFM slags, MSWI
bottom ashes, jarosite and paper residue. After the pre-extraction of
valuable components and the removal of heavy metals (i.e., fuming process
for NFM slags), the incorporation of the residual matrix in CAC-based
binders is evaluated in this study.
CAC is typically blended with calcium sulfate (C$) and PC in rapid-repair
building applications. The hydration products greatly depend upon the
ratio of these components in the mix. In general, the complete hydration of
a binary CAC-C$ system leads to an ettringite-rich binder given that
sufficient amount of sulfate is present. Whereas, the formation of
monosulfoaluminates is favored in low calcium sulfate systems. The phase
assemblage becomes more complex with the addition of PC where other
hydrates could form in greater proportions, such as C-S-H, hydrogarnet,
strätlingite, and monocarboaluminates. No previous study has been
published on the incorporation of residues similar to those of the
Among the residues investigated, the NFM slags are of particular interest
given their generally metal-depleted and amorphous nature. Previous
studies cited in this chapter, have demonstrated the reactivity of similar Fe-
rich slags in IP systems and of C4AF in cement systems. Using Mössbauer
spectroscopy, these studies consistently suggest the presence of several
oxidations states (2+, 3+) and coordination numbers (4, 5, 6) of Fe in both
systems. During polymerization or hydration, the oxidation process of Fe2+
to Fe3+ was discerned from the deconvolution of the Mössbauer spectra. Fe
uptake has been observed in several hydrate phases including siliceous
hydrogarnet, ettringite, monosulfate or monocarboaluminate, C-S-H and
even in several forms of Fe hydroxides.
Several techniques and methodologies have been employed to characterize
and even quantify slag reactivity. ATR-FTIR has been used in conjunction
with Mössbauer spectroscopy to indicate reactivity through distinctive
peak-shifts in the Si-O absorbance bands of Fe-rich slags. While the degree
of slag reaction in PC systems has been recently quantified using an
adapted EDTA-TEA-NaOH selective dissolution test. Finally, image
analyses from SEM-BSE micrographs and XCT 3D volumes also
demonstrated promising alternatives to quantify the degree of slag
reaction.
In the end, the challenge of this research work lies in the complexity of the
system. Reactions arising from the hydration of the ternary CAC, PC and
ASTOVEZA 2022 41
Chapter 1 ____________________________________________________
42 ASTOVEZA 2022
____________________________________________________ Chapter 2
Chapter 2
Materials, methods and residue
selection
Contents
2.1 Materials and formulation 45
2.1.1. The Socrates residues 45
2.1.2. Cements .. .. 48
2.1.3. Other blend components 48
2.1.4. Dry mix formulations 48
2.2 Methods ... . 50
2.2.1. Basic characterization of raw materials 50
2.2.2. Physical properties 53
2.2.3. Thermal and thermo gravimetric analyses 56
2.2.4. Tomography 56
2.2.5. Spectroscopy 59
2.2.6. Microscopy ... 64
2.2.7. Other techniques 69
2.3 Selection of industrial residue 69
2.3.1. Electric conductivity measurements 69
2.3.2. Heat Release 70
2.3.3. Setting time and mechanical strength 71
2.4 Evaluating jarosite residue as a sulfate source 73
Summary 79
ASTOVEZA 2022 43
____________________________________________________ Chapter 2
manuscript.
ASTOVEZA 2022 45
Chapter 2 ____________________________________________________
46 ASTOVEZA 2022
____________________________________________________ Chapter 2
Jarosite residue
Dried filter cake coming from a combined stream of a direct zinc sulphide
leaching stage and direct leaching reactors. This residue is often contained
in specially constructed tailing ponds or industrial landfills depending on
the levels of hazardous components.
ASTOVEZA 2022 47
Chapter 2 ____________________________________________________
2.1.2. Cements
Three types of cements were used in this study, namely:
Citric acid
Reagent grade admixture from Sigma-Aldrich, in granules/crystal form,
added to the dry-mix blends primarily as a set retarder.
48 ASTOVEZA 2022
____________________________________________________ Chapter 2
total CAC/HH
(in wt%) SCM/filler CAC HH
binder ratio
reference 0 82 18 100 4.5
quartz filler 24 62 14 100 4.5
granulated FS 30 57 13 100 4.5
slowly-cooled FS 29 58 13 100 4.5
fumed fayalite slag 29 58 13 100 4.5
bottom ash 24 62 14 100 4.5
In another case study under the screening test, the use of the jarosite residue
as a possible substitute to HH was explored. The binder composition used
is shown in Table 2.2. In this test, the variety of CAC used is Ciment
Fondu®.
Table 2.2 Binder composition for the case study on jarosite use as a sulfate source
in CAC- blended cement. Values are expressed in wt%
ASTOVEZA 2022 49
Chapter 2 ____________________________________________________
total CAC/HH
(in wt%) slag CAC HH TC PC
code binder ratio
A 4.5CAC-HH 30 57 13 0 0 100 4.5
B 1.6CAC-HH 30 43 27 0 0 100 1.6
A-TC 4.5CAC-HH-5TC 29 54 12 5 0 100 4.5
A-PC 4.5CAC-HH-30PC 30 33 7 0 30 100 4.5
B-PC 1.6CAC-HH-30PC 30 25 15 0 30 100 1.6
reference formulations:
REFcement (reference cement) no SCM nor quartz filler added
REFfiller (reference filler) slag replaced by equal volume of quartz
2.2. Methods
50 ASTOVEZA 2022
____________________________________________________ Chapter 2
(in wt%) CaO Al2O3 SiO2 FeO/Fe2O3 SO3 MgO K2O Na2O others
granulated FS 12.5 4.8 25.9 46.0 0.7 2.2 0.6 4.0 3.3
slowly-cooled FS 17.6 6.0 31.1 33.4 0.7 3.0 0.5 4.8 2.9
fumed fayalite slag 4.2 10.3 30.6 42.4 0.8 1.3 0.2 2.1 8.1
bottom ash 18.3 11.4 42.9 9.9 2.9 2.3 1.3 5.6 5.4
jarosite residue 4.3 1.0 6.4 22.0 51.2 0.5 0.3 4.8 9.4
paper residue 68.1 11.0 15.2 1.1 0.4 0.5 0.4 0.5 3.0
Secar 51® 37.7 52.3 4.9 1.8 0.2 0.4 0.0 0.0 2.7
Ciment Fondu® 37.8 40.3 4.2 14.6 0.1 0.4 0.1 0.1 2.5
PC 67.1 3.9 20.9 1.4 4.3 0.7 0.8 0.3 0.6
HH 41.7 0.1 0.8 0.1 56.5 0.1 0.0 0.0 0.7
quartz filler 0.0 0.0 99.8 0.0 0.0 0.0 0.0 0.0 0.2
ASTOVEZA 2022 51
Chapter 2 ____________________________________________________
Table 2.5 Minor components of some of the raw materials not detailed in Table 2.4
(in wt%) ZnO CuO Cr2O3 P2O5 PbO others
fumed fayalite slag 2.3 0.5 1.6 1.6 0.0 2.1
bottom ash 0.5 0.4 0.1 1.5 0.1 2.8
jarosite residue 3.3 0.1 0.0 0.1 4.4 1.6
In addition, the mineralogy of the raw materials was determined using the
D8 Advance X-ray Diffractometer (XRD) from Bruker coupled with
LynxEye detetor. CuK (1.54 Å) radiation was used with a step size of
-80°.
DiffracEVA and TOPAS software were used for the phase identification
and quantification. The amorphous content of the slag samples was
determined by applying the method described in et al., 2001 [145]
where 10 wt% of analytical grade zincite is added as an internal standard
during sample preparation. The XRD diffractograms and phase
quantification are detailed in Appendix B. Generally, both the granulated
Fe-rich slag and the fumed fayalite slag exhibited high amorphous content
(>90 wt%). The slowly cooled Fe-rich slag yielded a much lower
amorphous content (13 wt%), balanced by fayalite and spinel in the
crystalline fraction. On the contrary, the other residues (bottom ash,
jarosite and paper residue) were dominated by several crystalline phases.
The density of the raw materials shown in Table 2.6 was obtained through
helium displacement using the Micrometrics AccuPyc II 1340 pycnometer
system. Particle size distribution (PSD) presented along was determined
using the Malvern Mastersizer 3000 laser diffraction system. Specific
surface area (SSA) was quantified using both a semi-automatic Blaine Air
Permeability Apparatus and the Micrometrics BET Tristar 3000.
Table 2.6 Density, particle size distribution and specific surface area of the raw materials
density PSD SSA
Blaine BET
g/cm3 d10 d50 d90
(in wt%) (cm2/g) (m2/g)
granulated FS 3.67 1.1 7.9 50.5 4370 1.352
slowly-cooled FS 3.47 0.9 17.5 84.6 3200 1.216
fumed fayalite slag 3.50 1.6 13.7 61.5 3460 1.363
bottom ash 2.70 0.9 11.2 53.3 6160 3.865
jarosite residue 2.91 0.6 5.4 104.0 5253 4.706
paper residue 2.80 0.6 4.0 38.7 7730 8.133
Secar 51® 3.02 1.1 11.6 52.3 4270 1.019
Ciment Fondu® 3.23 1.6 16.4 75.8 3300 0.876
PC 3.17 0.9 8.5 25.3 3412 0.992
HH 2.64 0.6 3.7 22.7 7580 3.519
quartz filler 2.66 1.2 12.4 40.6 4580 1.118
52 ASTOVEZA 2022
____________________________________________________ Chapter 2
The goal of the grinding step during the sample preparation was to obtain
a comparable PSD and SSA between the SOCRATES residues and the
components of the blends to be substituted (CAC, PC and HH). Although
the PSD plot of the raw materials presented in Appendix C could support
that the curves relatively fall within a comparable range, it can be seen in
Table 2.6 that the bottom ash, jarosite and the paper residue have
significantly higher SSA. This is significantly related to the previously-
discussed difference in mineralogy of the materials and should be
considered during the screening test in section 2.3.
Setting time
Initial and final setting times were determined using the Vicatronic
Automatic Vicat Machine at 20°C (AFNOR room) without submersion to
water during curing. 1.3g of citric acid (0.3 wt% of the binder) was added
to formulations B, A-TC, A-PC and B-PC (refer to Table 2.3) to retard the
setting and therefore provide sufficient time for placing the freshly-mixed
blends in the moulds. Such low amount of citric acid (0.1-0.5 wt%) is not
expected to influence the phase assemblage evolution nor the strength
development of the mortars in ettringite-based binders according to
Nguyen et al., 2019 [82].
Mechanical strength
Compressive and flexural strengths were measured using the 3R Syntris
and the 3R RP 10/300 ELC press respectively, following the EN 196-1
standard. After demoulding, the mortars were cured under water at 20° C.
Each data point taken for a specific curing age represents the average value
from three mortar samples for flexural strength or six half samples for the
compressive strength with the standard deviation shown as error bars.
In parallel, the water-curing condition after extraction from the moulds was
substituted by air-curing conditions at 20°C and 65% relative humidity for
some of the slag-incorporating blends and their respective references in
order to investigate the effect of the curing condition to the mechanical
strength of the mortars. The results presented in Appendix D support the
advantage of following the water-curing procedure in discerning the
possible contribution of the slag to the long term (>28 days) strength
development of the mortars.
ASTOVEZA 2022 53
Chapter 2 ____________________________________________________
In this study, two criteria were used to evaluate the strength development
of the mortars calculated according to Equation 2.1 and 2.2. SAIcement
refers to the ratio of the mechanical strength of an SCM-containing mortar
to that obtained from its reference cement mortar. On the other hand,
SAIfiller is the ratio of the mechanical strength of the slag-containing mortar
to that obtained from its reference filler mortar. The evolution of these
indices over the curing period was used as a preliminary indication of slag
reactivity in Chapters 3 and 4.
Equation 2.1
Equation 2.2
Dimensional variation
Considering the expansive nature of ettringite binders, the dimensional
stability of the mortars was monitored during curing. Two methods were
used, described as follows:
Length variation was followed with the aid of gauge studs cast into
the opposite ends of the mortars. At different curing ages, the same
mortars were taken out of the water bath to record the length using a
digital length comparator apparatus (extensometer, see Figure 2.7)
54 ASTOVEZA 2022
____________________________________________________ Chapter 2
Figure 2.8 Walter+Bai shrinkage test set up (a) and the dimensions of the triangular
prism sample after the in-situ measurements of dimensional change (b, c).
Lengths are shown in centimeters.
ASTOVEZA 2022 55
Chapter 2 ____________________________________________________
Isothermal calorimetry
Ex-situ isothermal calorimetry was performed at 20°C on 10 g of freshly-
prepared paste samples using the TAM air (TA instruments) device.
where: bound water is expressed in grams (g) and mx represents the mass
of the sample at x = 50°C or 500°C.
2.2.4. Tomography
A methodology to estimate the degree of slag hydration using a
combination of X-ray computed tomography (XCT) and volume analysis
was introduced in this study. Micro-CT images of hardened slag-
containing paste samples were acquired using the Phoenix Nanotom S
56 ASTOVEZA 2022
____________________________________________________ Chapter 2
Figure 2.9 Picture of the XCT unit at the Materials Science Department of KU Leuven
used in this study (a) and of a hardened paste wrapped in Parafilm® and mounted in the
sample holder for XCT (b)
A 360° rotation scan using a step size of 0.15° was performed to obtain
tomographic projections. The X-ray source was regulated at 90 kV
(voltage) and 170 µA (current) in the nano focus mode 1. Tungsten on
CVD synthetic diamond was installed as a target and a 0.1mm thick copper
filter was mounted to harden the incident X-ray beams. A balance to
maximize the scan area while maintaining the precision lead to the
selection of 2.15 µm as the isotropic voxel resolution. This means that any
change in the scan properties below the particle size 2.15 µm was not
detected at this resolution. 2400 projections were taken with 500 ms
exposure time yielding a total scan time of 1 hour and 26 min per sample.
The position of the sample was saved for the next scan after a pre-
determined curing period.
Image reconstruction was performed in the Datos|x software producing
cross-sectional stack images. Using the 16-bit greyscale reconstructed
projections, threshold segmentation was done in the Bruker CT-Analyser
(CTAn) v 1.18.8.0+ software. A standard series of operations for image
correction described in Appendix E was followed in order to produce
binary image stack and calculate the total volume occupied by either of the
components (1 or 0) in the scan volume. This was then uploaded in the
ASTOVEZA 2022 57
Chapter 2 ____________________________________________________
2.2.5. Spectroscopy
57
Fe Mössbauer spectroscopy
This technique was used to elucidate the atomic environment of Fe, being
58 ASTOVEZA 2022
____________________________________________________ Chapter 2
the main component of the slag, and its evolution during the hydration.
57
Fe Mössbauer spectroscopy (WissEl GmbH) was operated using a
57
CoRh source and a metallic iron foil for the velocity scale calibration.
Although most of the spectra were measured at room temperature, some
samples were also analyzed at 77K to extricate overlapping signatures of
different Fe species during the fitting. The evaluation of the spectra was
performed by least-square minimization fitting of lines using the IMSG
software.
The Mössbauer spectra obtained from the raw slag is presented in Figure
2.10. Different Fe components were also fitted in this figure, using the
parameters presented in Table 2.7. More details on the model can be found
in Appendix F.
Figure 2.10 Deconvoluted 57Fe Mössbauer spectra of the slag (granulated FS)
measured at room temperature
Table 2.7 57Fe Mössbauer hyperfine parameters of the fitting proposed for the slag
shown in Figure 2.10. The different parameters are abbreviated as follows: isomer shift
(IS); central value of the quadruple splitting (QS); central value of the hyperfine magnetic
field (MF); total Gaussian-
hyperfine
The typical errors for the IS and QS is ± 0.02 mm/s and ± 5% for the AA
IS QS MF AA
Component (mm/s) (mm/s) (kOe) (mm/s or kOe) (%)
G1 1.14 1.91 0.00 0.42 30
G2 0.79 1.73 0.00 0.51 40
G3 0.38 1.04 0.00 0.17 9
M1 0.91 0.67 0.00 0.20 18
M2 -0.08 -0.01 330.93 14.65 3
ASTOVEZA 2022 59
Chapter 2 ____________________________________________________
total 100
The fitting demonstrates the diversity of Fe speciation in the slag despite
its seemingly simple crystallography composed of an amorphous phase (92
wt%). Components very similar to those found in Fe-rich slags in Siakati
et al., 2020 [117] and Peys et al., 2019 [116] were identified. Based on the
combination of Mössbauer parameters, the compositions were assigned as
follows: two doublets of Fe2+ (G1 and G2) and a doublet of Fe3+ (G3) in
the glass fraction; a doublet of Fe2+ (M1) and a sextet magnetic component
(M2) both in the crystalline fraction. The components were selected based
on the initial characterization of the slag (section 2.2.1). The reducing
environment in the furnace during the slag generation evidently resulted to
the dominance of the Fe2+ components in the system. The wide spread of
the quadruple s
incorporated in the amorphous fraction. The presence of two Fe2+
contributions in a similar slag has been previously reported to be a result
of the high irregularity of the Fe ions in the glass network [117] and
characterized by the asymmetric velocity resonant intensities. The higher
IS and QS values of G1 indicate the higher coordination number of Fe in
this component.
Furthermore, a minor Fe3+ component (G3) in the glass fraction was also
likely to be present in octahedral coordination. Wüstite was assigned as the
component M1 with the fitted area fixed according to the measured amount
of wüstite using XRD-Rietveld in the preceding section. On the contrary,
addition of the minor magnetic component (M2) improved the fitting of
the experimental spectrum despite the fact that no ferromagnetic minerals
were detected by using the other characterization techniques. Nonetheless,
it must be noted that the value could fall well within the measurement error
± 5% for AA in the fitting.
In addition, Figure 2.11 and Figure 2.12 show the spectra obtained for the
raw cement samples. Despite the indication from XRD (Appendix B) that
Fe in both of these materials is part of the C4AF crystal, a clear difference
in the spectra is obtained. That of the pure CAC has a similar pattern to the
spectra published in Honeybourne, 1980 [146] for a high alumina cement.
The primary contribution was attributed to a non-stoichiometric C4AF, but
also the presence of minor amounts of wüstite and magnetite was noted.
On the other hand, the spectra of the raw PC resembled that previously
obtained in Rai et al., 2009 [147] associated to the contribution of C4AF in
anhydrous PC. The noisy signals were caused by the significantly low Fe
content of these materials (< 2 wt% Fe2O3, Table 2.4) relative to the raw
slag. It follows that the signature of Fe that will be detected in the hydrated
pastes following the blends in section 2.1.4, can be assumed to almost
exclusively originate from the slag rather than the cement components.
60 ASTOVEZA 2022
____________________________________________________ Chapter 2
Figure 2.11 Deconvoluted 57Fe Mössbauer spectra of the CAC (Secar 51®)
measured at room temperature
Equation 2.5
ASTOVEZA 2022 61
Chapter 2 ____________________________________________________
62 ASTOVEZA 2022
____________________________________________________ Chapter 2
For the PC, two main peaks were identified positioned at 880 cm-1 and 520
cm-1 attributed to the multiple coinciding peaks of Si-O asymmetric
stretching vibration bands of C3S and C2S.
Finally for the HH, distinct S-O peaks at: 1150 and 1115 cm-1 (mode v3),
1006 cm-1 (mode v1), 653 and 593 cm-1 (mode v4); and O-H peaks at 1620
cm-1, 3660 cm-1 and 3550 cm-1 (mode v2, v3 and v1 respectively) were
identified.
ASTOVEZA 2022 63
Chapter 2 ____________________________________________________
2.2.6. Microscopy
Optical Microscope
An initial observation of the hardened paste samples under an optical
microscope (Nikon Eclipse LV150) revealed the distinct features of the
slag particles. Under polarized light (Figure 2.17 a), the dark color of the
raw slag particles makes them readily distinguishable from the other
components present such that a simple image analysis using ImageJ can be
used to reconstruct binary images as demonstrated in Figure 2.17 b.
Following the concept and assumptions described in section 2.2.4, a
similar methodology for calculating the degree of slag hydration can be
performed based on the change in the total slag volume over time.
However, the main drawback in using images from the optical microscope
is the limited resolution. In order to address this issue, the use of SEM-
BSE micrographs was opted in this study. Nonetheless, these initial
observations using the optical microscope could be useful for a rapid
evaluation for future studies.
64 ASTOVEZA 2022
____________________________________________________ Chapter 2
In Figure 2.18, images from optical microscope are presented side by side
with that of the SEM in BSE mode. Both techniques revealed the distinct
dendritic surface of the raw slag particles, a notable feature rendered by
the Fe-O dendrites in granulated slags [148]. This facilitated the
identification of the slag particles under the microscope in this study and
consequently aided the selection of image analysis operations for the
quantification of the degree of slag hydration.
Equation 2.6
where: is degree of hydration; A is the total area of the slag. In this case,
the degree of slag hydration is estimated after 90 days of hydration.
A comparison of the values of the degree of slag hydration obtained using
image analysis of SEM-BSE micrographs versus that obtained from the
volume analysis of tomography is presented in Chapter 3.
ASTOVEZA 2022 65
Chapter 2 ____________________________________________________
Figure 2.19 TEM images of pure synthetic C-S-H (a); C-S-H with low, 0.010-0.014
mmol/g, Fe uptake; (b) and C-S-H with high, 0.039-0.048 mmol/g, Fe uptake
modified from Siramanont et al., 2021 [133].
66 ASTOVEZA 2022
____________________________________________________ Chapter 2
Figure 2.20 STEM-EDS Fe-maps of C-S-H foils possibly with Fe incorporated in the
structure (a); and C-S-H foils intermixed with Fe-rich spherical nanoparticles,
possibly ferrihydrites or hydrogarnet. The images were modified from
Siramanont et al., 2021 [133].
ASTOVEZA 2022 67
Chapter 2 ____________________________________________________
Electric conductivity
Electrical conductivity measurements were made using the Mettler-Toledo
SevenMulti for 0.1 g/L and 0.5 g/L suspensions with no pH adjustment.
The temperature was controlled at 20°C using a 1 L thermostatic beaker
filled with 900 mL of deionized water (pH 7.1) where the solid is
introduced. Measurements were taken at 5 s interval and the suspension
was constantly agitated at 500 rpm. The beaker was sealed using
Parafilm® to limit atmospheric carbonation. A schematic diagram of the
set-up adapted from Maach, 2016 [154] is presented in Figure 2.22.
ICP-AES
5 mL solution was extracted after 30 min of conductivity test using a
syringe equipped with 0.2 µm polyethersulfone filter. The filtrate was
stabilized for ICP-AES analysis by adding 5 vol% nitric acid (67%,
NORMATOM®).
68 ASTOVEZA 2022
____________________________________________________ Chapter 2
ASTOVEZA 2022 69
Chapter 2 ____________________________________________________
the reference binder. In Figure 2.23 b, a closer look on the trend shows
that although the overall degree of conductivity is closely comparable
among these other residues, the kinetics appear to have a particular pattern
for the highly amorphous slags. Their conductivities continue to increase
throughout the first 30 min unlike the other residues which appear to
stabilize even within the first 2 min of the test.
Figure 2.23 Conductivity measurements on raw materials using 0.5 g/L suspension
in deionized water at 20°C
70 ASTOVEZA 2022
____________________________________________________ Chapter 2
Figure 2.24 (a) Heat flow and (b) cumulative heat release of paste samples for the
screening test
On the other hand, the cumulative heat release graph shown in Figure 2.24
b demonstrates the overall decrease in heat release with the addition of the
SOCRATES residues. Only the blend incorporating bottom ash has
achieved a comparable level of heat release as the reference formulation
after more than 90 hours of hydration. This could be related to the
relatively high specific surface area (Table 2.6) of the bottom ash, possibly
providing more nucleation sites for hydration. For the other residues, the
differences in heat values may not be sufficient enough to draw conclusive
comparisons.
Figure 2.25 Initial and final setting time of mixtures for the screening test
The addition of the residues generally, but not substantially, delayed the
setting compared to the reference cement formulation. Only the blend
ASTOVEZA 2022 71
Chapter 2 ____________________________________________________
Figure 2.26 (a) Compressive strength evolution and (b) strength activity index (SAI) of
standard mortars for the screening test
Considering the results from the raw material characterization along with
the properties of blends presented in this section, the highly amorphous
slags were determined to have the highest potential as SCM (at 30 wt%
addition) in the CAC-rich system. Consequently, Chapters 2 and 3 will be
focusing on the properties and hydration mechanism of blends
incorporating the granulated FS (slag). In addition, the challenge in
improving the early age performance of these blends by activating the slag
will also be addressed in the next chapters.
72 ASTOVEZA 2022
____________________________________________________ Chapter 2
Figure 2.27 Compressive strength evolution of blends with jarosite residue addition
Based on the XRD patterns of the hydrated pastes (Figure 2.28) after 28
days of curing, the phase assemblage of the reference blend is dominated
by ettringite and gibbsite with small amount of monosulfoaluminate. As
the jarosite residue is added in higher quantities, the peaks associated to
ettringite diminish. However, the patterns generally remained consistent
until the HH was completely substituted by the residue (100JAR). At
100 wt% substitution, the peaks associated to the raw jarosite residue
(Appendix B, Figure B.5) particularly those of jarosite, gypsum and quartz
became more apparent. The height of the main ettringite peak detected at
and appeared to be slightly shifted
to 9.12°. As mentioned in Chapter 1, these shifts are often attributed to Al-
Fe solid substitution [131] which could be a possibility considering the
ASTOVEZA 2022 73
Chapter 2 ____________________________________________________
Figure 2.28 XRD patterns of hydrated pastes with different substitution level of
jarosite residue after 28 days of curing. The phases are abbreviated as follows:
(Ett) ettringite, (Ms) monosulfoaluminate, (Gib) gibbsite, (CA) monocalcium
aluminate, (Jar) jarosite, (Gyp) gypsum, (Sul) elemental sulfur, and (Qua) quartz.
Only the principal diffractogram peaks are labelled.
74 ASTOVEZA 2022
____________________________________________________ Chapter 2
Comparing the SEM images of the paste samples at 0 wt% (a) and 100 wt%
(b) substitution levels (Figure 2.30), ettringite needles are observed to be
less abundant as HH is replaced by the jarosite residue.
Figure 2.30 SEM-SE images of (a) no jarosite and (b) 100JAR paste samples
after 3 days of curing. Encircled in red are needle-like structures generally associated to
ettringite.
It can be inferred that the lower amount of sulfates available in the solution
lead to decreased ettringite formation an observation consistent with the
phase assemblage evolution obtained through XRD. As previously
mentioned in Chapter 1, maximizing ettringite formation is often desirable
in sulfated CAC systems due to its strength-giving properties [9,156]
which could be associated, although not directly correlated, with the
obtained strength values of the blends containing the jarosite residue.
ASTOVEZA 2022 75
Chapter 2 ____________________________________________________
Other ions (Na, Al, Si, K, Mg, P and Fe) were also released from the
jarosite residue during the dissolution test quantified in Figure 2.31. One
of the biggest concerns in the valorization of the jarosite residue as SCM
is the possible release of toxic elements and heavy metals. Although these
components were not analyzed in ICP for this study, the dissolution of the
other ions could signal that the jarosite residue is indeed reactive in this
system and therefore the toxic components should be monitored more
closely. For instance, the sulfur content of the raw residue alone is already
alarming (20-30 wt% see Appendix B, Figure B.5). It is clear that this
residue is not the typical jarosite by-product previously reported to have
potential use as raw meal substitute to gypsum [157]; or as mineral additive
in non-structural concrete [158]. Going back to the source of this by-
product (section 2.2.1), the complexity of the composition of this residue
can be traced from combining the streams of different leaching processes
into one filtration unit. In fact, the sulfur content can be totally avoided by
dedicating separate filtration lines for each reactors. This solution, of
course, may not consider the other practical aspects of the production.
Figure 2.31 Comparison of the ion concentrations after the dissolution test
measured using ICP-AES
76 ASTOVEZA 2022
Chapter 2 ____________________________________________________
Summary
This chapter details the materials and methods used in this study. Residues
coming from a single production batch were used throughout this study to
maintain uniform results. They were homogenized, sampled, dried and
grinded to obtain a PSD as close as possible to that measured for the CAC.
The characterization of the residues revealed a diverse mineralogy related
to their origin. Their main properties are highlighted below:
Granulated FS
An iron-rich highly (92 wt%) amorphous slag originating from lead-zinc
production. Mössbauer spectroscopy revealed that the Fe speciation in slag
is dominated by Fe2+. The Si-O bonds in this material were detected using
ATR-FTIR. When added in CAC-rich blended cement, the dark color, high
density and Fe content of the unreacted slag particles could facilitate its
differentiation from the rest of the cement matrix.
Slowly-cooled FS
An iron-rich slag with the same origin as the granulated FS but more
crystalline in nature (only 13 wt% amorphous) due to the slower cooling
rate employed. Its composition is dominated by fayalite and spinel.
Fumed fayalite slag
An iron-rich highly (94 wt%) amorphous slag originating from secondary
copper production. Although the basic properties are closely comparable
to those of the granulated FS, this batch of material still contains 2.3 wt%
of ZnO.
Bottom ash
A siliceous by-product of a municipal waste incinerator with a highly
heterogenous composition mainly consisting of quartz, calcite, muscovite
and wollastonite. Fragments of unburnt components (i.e., plastics, glass)
were observed in the sample received. Its SSA was much higher than those
of the slags in this study despite displaying a similar PSD.
Jarosite residue
A sulfate-rich (51 wt% SO3) by-product of zinc leaching. Its mineralogy is
dominated by jarosite, gypsum and sulfur. Its substantial Fe content
(22 wt% Fe2O3) is linked to the jarosite component. Significant traces of
Pb and Zn were also detected. Similar to the bottom ash, high SSA value
was also obtained.
Paper residue
A lime-rich (68 wt% CaO) by-product of paper recycling. No further
grinding was executed due to its very fine nature as-received. The electric
78 ASTOVEZA 2022
____________________________________________________ Chapter 2
ASTOVEZA 2022 79
____________________________________________________ Chapter 3
Chapter 3
Fe-rich slag addition in binary CAC-
HH binder
Contents
Introduction . 83
Conclusion 135
ASTOVEZA 2022 81
____________________________________________________ Chapter 3
Introduction
In this chapter, the addition of the highly amorphous Fe-rich non-ferrous
metallurgy (NFM) slag in calcium aluminate cement (CAC)-calcium
sulfate hemihydrate (HH) binder systems was investigated. In Chapter 2,
this slag was identified to have the best potential as a supplementary
cementitious material (SCM). The slag displayed contribution to the long
term (>28 days) strength development of the standard mortars. In addition,
its leachates (CuO and ZnO) were determined to be both below the
detection limit, in contrary to the 2.3 wt% ZnO measured in a similar
granulated slag among the selection.
The properties of formulations, A, B, and A-TC (Chapter 2, Table 2.3)
were examined in this chapter. All of these formulations contained
30 wt% of slag. Formulations A and B have CAC/HH ratio of 4.5 and
1.6, respectively. On the other hand, formulation A-TC has CAC/HH ratio
of 4.5 similar to that of the formulation A, but with 5 wt% addition of the
paper residue (TC). This lime-rich (68 wt% CaO) material is highly soluble
in water and is already being commercially marketed as a mineral additive
for cement and concrete. Considering these properties, it has been
evaluated separately as a possible activator in formulation A, aiming to
induce slag dissolution and contribute Ca2+ ions during the early age
(<28 days).
Several techniques were employed to characterize the properties of the
blends for up to 1 year. Mechanical strength and dimensional stability were
monitored using mortar samples. The degree of slag dissolution was
estimated using SEM-BSE, X-ray Computed Tomography (XCT) and
Mössbauer spectroscopy. The phase assemblage of the binary systems
were simulated using thermodynamic modelling in function of the CAC
proportion. Experimentally, several techniques, including isothermal
calorimetry, XRD, TGA, and ATR-FTIR were used to follow the evolution
of the phase assemblage using paste samples. Finally, the fate of the Fe
from the slag was described based on the results of Mössbauer
spectroscopy, TEM and scanning transmission electron microscope
(STEM).
The methodology for the characterization techniques was specifically
adapted to the distinct properties of the slag primarily related to its high Fe
content (i.e., high density, high atomic mass, Mössbauer effect). For this
reason, the CAC used was changed from Ciment Fondu® in the screening
test to Secar 51®. The latter contains less Fe2O3 (< 2 wt%) and was
therefore expected to interfere less with the signals of the Fe from the slag.
ASTOVEZA 2022 83
Chapter 3 ____________________________________________________
Figure 3.1 Initial and final setting time of formulation A and its references
On the other hand, the reference filler formulation displayed a more
significant setting retardation despite following an equal volume
substitution as the slag. Considering that the quartz filler and the slag have
a closely comparable PSD and SSA (Chapter 2, Table 2.6), this difference
84 ASTOVEZA 2022
____________________________________________________ Chapter 3
The compressive (Figure 3.2) and flexural (Figure 3.3) strength evolution
of formulation A and its references were followed using standard mortars
over 1 year of curing under water. The evolution of the SAIcement values are
presented in the secondary axes (right).
Figure 3.2 Compressive strength of formulation A (primary axis) and its SAI evolution
(secondary axis)
Figure 3.3 Flexural strength of formulation A (primary axis) and its SAI evolution
(secondary axis)
Similar to the observed trend during the screening test in Chapter 2, the
addition of the slag also clearly resulted in the overall decrease in strength
compared to the reference cement formulation. The same effect has been
previously reported in PC systems during the early hydration stage (<28
days) of a similar NFM slag [160] and as well as other SCM [8]. Although
this trend was less evident in terms of the flexural strength, the decrease in
the compressive strength reached as much as 31% in the first day of curing
with SAI value 0.64. It was not until after more than 28 days that the
compressive strengths of formulation A and its reference became
comparable at 39 MPa versus 41 MPa, respectively. This was exhibited by
the abrupt increase in the SAI values from 28 to 90 days. This could be
ASTOVEZA 2022 85
Chapter 3 ____________________________________________________
Figure 3.4 Length variation of formulation A and its reference obtained from Walter+Bai
shrinkage test for the first 24 hours of hydration. Measurements were continued using an
extensometer for the demoulded samples cured under water.
It can be seen that the addition of the slag did not significantly alter the
dimensional stability of the blend. An overall shrinkage was observed
during the first 24 h for both formulations. The mechanism of ettringite
formation was hypothesized to slightly alter trend, such
that shrinkage was counteracted in the reference cement formulation by the
fast formation of interlocking ettringite from 6 to 18 h. On the contrary,
the slag-containing blend displayed a more abrupt shrinkage at this period,
which could be an indication of less amount of expansive hydrates formed.
86 ASTOVEZA 2022
____________________________________________________ Chapter 3
Beyond 24 hours, curing under water triggered a slight expansion for both
formulations. Not only was additional water provided to stimulate
hydration in this step, but also drying shrinkage was avoided throughout
the remaining duration of curing.
The overall dimensional stability of formulation A was reflected through
the progressive strength development in mortar samples. The addition of
the slag even appeared to have positively contributed to the stability
keeping the values closer to zero in Figure 3.4. In the end, it should be
considered that despite this observation, the scale of the length
measurements prese
within the confidence interval of the test and consequently may not provide
conclusive quantitative comparison.
ASTOVEZA 2022 87
Chapter 3 ____________________________________________________
The strength values in this high calcium sulfate system were considerably
higher than those obtained from the low calcium sulfate formulation, B.
However, the SAI values were generally lower implying that the observed
increase in strength was more likely associated to the hydration of CAC
and HH, producing more ettringite, rather than an increased rate or extent
of slag hydration. An indication of this was the observed expansion of
mortars in Figure 3.7. It will also be confirmed from the techniques in the
following section.
88 ASTOVEZA 2022
____________________________________________________ Chapter 3
3.2.1. SEM-BSE
The SEM was initially used in this study to characterize the microstructure
of the blends. Figure 3.8 (a) shows an SEM-BSE micrograph of
formulation A mortar sample after 90 days of hydration. The slag particles
appear the brightest due to their high Fe content (high atomic number), and
their boundaries remained intact with no apparent formation of hydration
rim around the surface of the particles. On the contrary, the hydration of
CAC particles was more evident with their dissolved surfaces gradually
blending into the matrix. The lower reactivity of the slag versus the other
binder components (CAC and HH) was further supported by the EDS maps
(Figure 3.8 b-f) revealing the massive dispersion of Ca, Al and S
ASTOVEZA 2022 89
Chapter 3 ____________________________________________________
throughout the interstices, while Fe and Si are localized in the slag and
sand particles, respectively.
Figure 3.8 (a) SEM-BSE micrograph of formulation A mortar after 90 days of hydration
along with (b-f) the corresponding EDS maps for Fe, Ca, Al, S and Si
90 ASTOVEZA 2022
____________________________________________________ Chapter 3
Figure 3.10 SEM-BSE micrograph of formulation A pastes after (a) 1 day and (b) 90 days
and their corresponding binary images obtained using Image J software
Based on the image analysis, the total slag area and degree of hydration
after 90 days of curing were estimated. The results are presented in Table
3.1. Despite the abrupt increase in SAI values (Figure 3.2) after 90 days
of hydration, it turns out that only about 14 % of the slag has reacted at this
point. Otherwise, this could also signify that the properties of paste and
mortar samples at any given curing time were not directly comparable,
particularly considering the difference in their interfacial transition zones
[9] resulting from the w/b ratios used.
ASTOVEZA 2022 91
Chapter 3 ____________________________________________________
Table 3.1 Estimated slag area and hydration degree using SEM-BSE + image analysis
curing time total slag area degree of slag hydration
(days) (µm2) (%)
1 5.63 0
90 4.83 14.2
Although the huge compositional difference between the slag and the rest
of the components of the hydrated paste greatly favored this methodology,
certain assumptions and challenges encountered must be carefully
considered. For instance, the highly contrasting abrasion resistance of the
blend components (i.e., slag versus CAC) resulted in uneven surfaces
during polishing. The hydrate products and the unreacted CAC were more
easily scratched compared to the unreacted slag particles leaving the latter
embossed on the surface. This was manifested in the micrographs as
shadows around the slag particles which could significantly influence the
image analysis. Another important issue was representativeness. The need
to prepare polished sections required the use of separate samples to be
analyzed per curing time. This means that we were not actually following
the dissolution of the same slag particles over time, rather we were simply
assuming that the total areas detected were representative of each curing
time. Considering that we were working on the micro-scale, the randomly
selected micrographs might not accurately represent the overall slag
distribution. In fact, areas where slag particles appeared to be agglomerated
and areas with particularly low amount of slag were observed as shown in
Figure 3.10 (b) 3 and 4, respectively.
In order to address the above-mentioned challenges, a non-destructive
technique using a combination of XCT and volume analysis was discussed
in the next section. Although the methodology revolved around the similar
principle (following the changes in the slag area/ volume over time), the
need for preparing polished sections and using different samples per curing
time was eliminated.
3.2.2. Tomography
The degree of slag hydration of formulations A and B were followed after
1, 28, 90, 180 and 365 days using a single hardened paste sample per
formulation. The threshold segmentation to select the slag particles was
optimized, guided by the observations from the SEM (EDS point ID). The
major steps involved were illustrated in Figure 3.11. In this figure, a red
arrow is drawn in the 2D reconstructed images (a). The plot profiles (b)
corresponding to the pixels under the arrows demonstrate that the selection
of the lower threshold value for the slag particles was greatly facilitated by
the distinct difference in the range of grayscale index between the slag and
the rest of the components. The binary images (c) were obtained following
92 ASTOVEZA 2022
____________________________________________________ Chapter 3
Figure 3.11 Illustration of the steps involved in the volume analysis of XCT scans
There were two important details to consider in analyzing the XCT results:
1. Variations in the scanned VOI was inevitable despite using the same
paste samples over time mainly due to the dimensional changes
during hydration. For instance, the particles in the edges could move
in (shrinkage) or out (expansion) of the VOI.
2. The voxel resolution of the scans was set to 2.15 µm in order to
cover a sufficiently representative volume while optimizing the scan
time. This means however that the slag particles with a diameter
ASTOVEZA 2022 93
Chapter 3 ____________________________________________________
curing (<90 days) was not significant enough to contribute to the strength
evolution, regardless of the CAC/HH ratio.
Figure 3.12 Estimated slag volume and hydration degree using XCT + volume analysis
Considering that only 40-50 % of the slag has reacted even after 1 year,
the interest in exploring the reactivity of slag in modified systems was later
94 ASTOVEZA 2022
____________________________________________________ Chapter 3
addressed in this study with the addition of the lime-rich paper residue
(Chapter 3, section 3.5) and PC (Chapter 4).
Despite the concern on the risk of underestimation due to the voxel
resolution, the estimated slag hydration degree (27 %) still ended up higher
than that (14 %) from the image analysis using SEM micrographs.
Although the two techniques were not directly comparable (i.e., volume-
versus area-based calculation; number of images processed), these findings
still open questions on the sensitivity of these image-based calculations.
Sensitivity Analysis
The influence of the grayscale threshold segmentation and the use of
different samples (hydration stopped) per curing time were evaluated to
test the sensitivity of the XCT methodology employed. These were among
of the slag particles) during the scans compared to the rest of the
components (image contrast).
For the second factor, samples of formulation A hardened pastes were
scanned after stopping the hydration at 1 and 90 days (one sample per
curing time). The VOI used for volume analyses was also changed to
cylinder (Figure 3.13) as opposed to the previously-presented cubic
samples. Smaller volumes (9 mm3 for the cubic VOI and 6.28 mm3 for the
cylindrical VOI) were scanned and reconstructed.
ASTOVEZA 2022 95
Chapter 3 ____________________________________________________
observation that the slag particles were easily distinguishable from the rest
of the components both in the SEM and XCT images.
Figure 3.14 Estimated (a) slag total volume and (b) slag hydration degree
from the sensitivity analysis for formulation A
On the other hand, the use of different samples per curing time and a
smaller VOI resulted in -0.43% volume or -5% slag hydration degree
difference. This highlights the advantage of using XCT over SEM, as a
rapid non-destructive test that could autonomously take a large sample size
while minimizing factors that could impose human bias.
In the end, selecting the more suitable methodology is a matter of
balancing several factors to satisfy the goal of the study. The focus must
be directed towards the rate rather than the absolute values of the slag
hydration to avoid misinterpretations. In any case, quantification must be
supported by complementary techniques (i.e., selective dissolution)
especially if determining the absolute values is crucial.
96 ASTOVEZA 2022
____________________________________________________ Chapter 3
condition and were calculated at infinite curing time (until stability). Other
assumptions along with the codes used to generate these models are
detailed in Appendix I. The models were calculated in function of the CAC
content (x-axis), which could be interpreted in relation to the CAC/HH
ratios (1.6 and 4.5) chosen in this study.
In Figure 3.15, it can be seen that gypsum is formed with the excessive
HH at lower (< 53%) CAC percentages. On the opposite extremity,
hydrogarnet is formed from the hydration of the remaining pure CAC when
sulfates are depleted. As the model does not consider the kinetics of
hydration, the formation of metastable calcium aluminate hydrates
mentioned in Chapter 1 is experimentally expected, prior to their ultimate
conversion to hydrogarnet at longer curing ages. Moreover, the phase
assemblage of reference formulations A and B are shown to be dominated
by monosulfoaluminate and ettringite, respectively. This model better
illustrates the interest in the CAC/HH ratios selected in this study. These
formulations both avoid regions of excessive sulfates marked by gypsum
formation, which could have detrimental effects (i.e., cracking, spalling)
related to secondary ettringite formation [162].
In both reference formulations, the silicates from the CAC are expected to
be incorporated in strätlingite, while the Fe from C4AF partakes in
siliceous hydrogarnet formation. It can also be seen that formulation B
dwells within the region of maximum hydrate formation by mass, such that
the 65 g of water added to the 100 g of anhydrous blend is barely sufficient
for full hydration.
Changes in the phase assemblage are evident with the addition of 30 wt%
slag in the system (Figure 3.16). As indicated in Appendix I, only 50 vol%
of the amorphous fraction of the slag was assumed to react during the
modelling, based on the XCT estimations after 1 year of hydration. The
black region, along with the minor (<1 at%) components of CAC and HH.
Fe appears to be incorporated in ferrihydrite at lower CAC percentages.
The dominant Fe-bearing phase at higher percentages is still Fe-siliceous
hydrogarnet represented in the GEMS in several phases as follows:
C3FS0.84H4.32
C3FS1.34H3.32
C3(A,F)S0.84H (solid-solution between Al and Fe)
ASTOVEZA 2022 97
Chapter 3 ____________________________________________________
Figure 3.15 Phase assemblage of the reference (no slag) binary binder calculated from
GEMS thermodynamic modelling software. CAC/HH content is varied in the x-axis
Figure 3.16 Phase assemblage of the binary binder incorporating 30% slag with varying
CAC and HH proportions, calculated from GEMS thermodynamic modelling software
98 ASTOVEZA 2022
____________________________________________________ Chapter 3
ASTOVEZA 2022 99
Chapter 3 ____________________________________________________
Figure 3.19 XRD patterns of formulation A (solid black lines) and its reference
measured over 1 year of curing. Phases are abbreviated as follows: (Str) strätlingite,
(Ett) ettringite, (Ms) monosulfoaluminate, (Geh) gehlenite, (Hdg) hydrogarnet and (Wue)
wüstite
Unlike in the slag-containing formulation, hydrogarnet was detected after
1 year of curing in the reference formulation. It has been demonstrated
[163,164] that the formation of hydrogarnet can destabilize several AFm
phases in low-sulfate systems below 50°C. Based on thermodynamic
modelling [164], AFm only becomes stable when hydrogarnet is
suppressed under the assumption that its formation is not
thermodynamically-favored in the early age of hydration. This instability
could be one of the reasons behind the increasing peak intensities from 90
days to 1 year giving the impression of ettringite re-formation observed in
the reference formulation.
Strätlingite peaks were consistently present from 7 days to 1 year in the
slag-containing formulation which could have contributed to its strength
development, supporting previous studies [165,166] where the strength-
giving property of strätlingite has been demonstrated. In general, the
addition of siliceous SCM inducing strätlingite formation has been
associated to denser pore structures [165,167]. At the same time, the
preferential formation of hydrates (i.e., ettringite, strätlingite) other than
the metastable calcium aluminate phases (i.e., CAH10, C2AH8) with the
addition of SCM and calcium sulfates inhibits the conversion phenomenon
[168] typically considered undesirable in pure CAC systems. The acidic
nature of the Fe-rich slag could also slow down the conversion reactions
[169]. On the contrary, the appearance of strätlingite peaks in the reference
formulation occurred later between 7 and 28 days, diminishing again with
the formation of hydrogarnet and ettringite from 180 days.
Moreover, gehlenite appeared to be consumed faster in the reference
formulation with peaks greatly reduced beyond 7 days. This was not the
case for the slag-containing formulation likely due to the additional
silicates contributed by the slag in the system. On the other hand, wüstite
from the slag seemed to be unreactive, displaying consistent peak
intensities throughout the hydration period. This agrees with the findings
from a previous study [117] suggesting the preferential dissolution of the
amorphous component of similar slags over that of its crystalline
component.
Contrary to the more complex phase assemblage evolution of formulation
A and its reference, less changes were observed between the XRD patterns
of formulation B and its reference shown in Figure 3.20. The high-sulfate
system was indeed dominated by ettringite, stable throughout the hydration
period. This result agrees with the initial findings based on the strength
development and heat release. For both formulations, however, no new Fe-
containing hydrate was detected based on the XRD patterns alone. Despite
previous studies [131,170] suggesting the solid substitution of Fe to Al in
hydrates such as ettringite, monosulfoaluminate or siliceous hydrogarnet,
there was no clear peak shifts in the diffractograms presented that could
established at this point based on the XCT results, this could be an initial
indication that the Fe could be incorporated instead in an amorphous or in
a nano-crystalline hydrated phase. In order to unravel the fate of the Fe in
this system, further complementary techniques were employed.
Figure 3.20 XRD patterns of formulation B (solid blue lines) and its reference
measured over 1 year of curing. Phases are abbreviated as follows: (Ett) ettringite,
(AFm) AFm, (Ms) monosulfoaluminate, (Gib) gibbsite, (Geh) gehlenite, (Hdg)
hydrogarnet, (Gyp) gypsum and (Wue) wüstite
Figure 3.21 Evolution of the DTG curves of (a) formulation A and (b) formulation B
along with their corresponding references. Slag-containing formulations are shown in
solid lines. Phases are abbreviated as follows: (Ett) ettringite, (AFm) AFm, (Str)
strätlingite, (Gib) gibbsite, (Hdg) hydrogarnet
Figure 3.22 Total bound water content of formulations A, B and their references
Figure 3.23 ATR-FTIR spectra of formulation A (solid black lines) and its reference.
identified.
Compared to the inorganic polymers investigated in Siakati et al.,
2020 [117], the cement systems in this study displayed a more complex
spectra due to the co-existence of various crystalline and amorphous
phases. The general phase assemblages of the two formulations were
Figure 3.24 ATR-FTIR spectra of formulation B (solid blue lines) and its reference.
identified
3.4.1. TEM
The TEM images of formulation A paste after 1 day of hydration are
presented in Figure 3.25. The observed phases were consistent with the
assemblage described in the preceding section. Several individual rods of
ettringite (A2 and A6); plate of anhydrous CA (A4); Al-rich gels (A1, A5,
A7 and A8) similar to gibbsite; and intermixed phases (i.e., A3) were
observed. No significant amount of Fe was detected in any hydrated phase
at this stage of hydration. Moreover, the amount of S detected through EDS
analysis in rod/needle-like phases labelled as ettringite was significantly
lower compared to what could be theoretically expected. This discrepancy
is linked to the uncertainties associated to the EDS analysis particularly for
S in ettringite, discussed in section 2.2.6.
Figure 3.25 TEM images and the atomic composition of marked areas in the (A)
4.5CAC- HH paste sample after 1 day of hydration
At longer curing ages (>28 days) when the degree of slag hydration became
more significant, Fe was detected in the Al-rich gels with a slightly higher
amount of Si compared to the gels observed in the early days of hydration.
Figure 3.26 shows the general morphologies (a) and the Fe-containing
intermixed phases (b-c) observed in formulation A. At 90 days, more plates
108 ASTOVEZA 2022
____________________________________________________ Chapter 3
Figure 3.26 TEM images and the atomic composition of marked areas in the (A)
4.5CAC-HH paste sample after long term curing
For some of the Fe-bearing gel phases (i.e., A12), the Fe signal was notably
higher than those observed in other gels (i.e., A9 and A11). To better
understand their nature, the diffraction mode (Figure 3.27) of TEM was
used to characterize gels at (a) 180 days and (b) 1 year, both containing
diffraction spots compared to the A13 gel at 180 days. The former could
be interpreted to have some crystalline component/s [175] intermixed with
gel phase/s; while the latter was clearly more poorly-structured [176]. The
possibility of having erroneously analyzed unreacted slag particles as part
of the gels was excluded, not only due to the spot size (only a few hundred
nanometers) but also due to the sensitivity (Figure 3.28) of these gels.
Even the gels at 1 year (A16 and A17) were immediately destroyed after
the EDS analysis, suggesting that they were, most likely, hydrated phases.
Figure 3.27 TEM images, diffraction patterns and the atomic composition of Fe-
containing gels in the (A) 4.5CAC-HH paste sample
Figure 3.28 TEM images before and after EDS analysis of Fe-containing gels in the (A)
4.5CAC-HH paste sample
Figure 3.29 TEM images and the atomic composition of crystalline phases in the (A)
4.5CAC-HH paste sample
Figure 3.30 TEM images and the atomic composition of marked areas in the (B) 1.6CAC-
HH paste sample
3.4.2. STEM
Although TEM facilitated the identification of individual phases with
relatively less intermixing compared to the previous attempt using SEM
(section 3.2.1), the homogeneity of the Fe-bearing phase observed in TEM
was still in question. STEM was therefore employed to create elemental
maps focusing on these phases.
Another area of the same sample was examined and presented in Figure
3.32. The Fe content detected for the amorphous hydrate gels (A27 and
A28) remained below 12%, consistent with the general trend observed
from the TEM and STEM analyses. In addition, unreacted slag particles
(i.e., A26, A29) were detected. The EDS analysis (A26) and diffraction
pattern (EDS-A29) suggest that these particles could be wüstite rather than
the amorphous fraction of the raw slag, supporting the earlier observations
that the mineral fraction of the slag was less (if at all) reactive during 1
year of hydration.
Figure 3.31 STEM elemental mapping, diffraction patterns and atomic composition
of marked areas in the (A) 4.5CAC-HH paste sample after 180 days of curing
Figure 3.32 STEM elemental mapping, diffraction pattern and atomic composition
of marked areas in the (A) 4.5CAC-HH paste sample after 180 days of curing
The ternary diagram in Figure 3.33 shows the distinct Al-rich cluster
occupied by the Fe-bearing hydrate gels. In view of the limited number
points, however, no definite conclusion could be drawn to compare the
variation in the composition between the gels in formulation A or B, nor
among the samples at different curing ages. Reference points for gibbsite,
strätlingite, katoite and Si-hydrogarnet were also added in the diagram
based on their empirical formula. The composition of the hydrate gels was
observed to fall between that of gibbsite and stätlingite/katoite, but
remained distinct from these pure phases.
Further attempts were made using STEM after 1 year of hydration to
identify the Fe-containing phase which exemplified some diffraction spots
in previous analyses (i.e., EDS-A24, Figure 3.31). Fe maps and diffraction
patterns of the Fe-bearing intermixed phases observed in formulation A are
presented in Figure 3.34 and Figure 3.35. The inhomogeneity of Fe
distribution is clear from the maps, consistent with the observation from
180 days. A distinct rod-like to lamellar morphology was also observed for
the phase with higher Fe content (i.e., A30 and A31, Figure 3.34) in the
updated 1 year sample. In an attempt to measure the interatomic spacing
(d-spacing) for the identification of this phase, a more focused STEM
image was taken as shown in Figure 3.36.
Figure 3.35 STEM image focused on an Fe-rich region (a); along with its corresponding
annular dark field image (b); Fe map from EDS analysis (c); and the diffraction pattern
of marked areas from the (A) 4.5CAC-HH paste sample after 1 year of curing
Figure 3.36 STEM images focused on the needle-like Fe-rich phase (a, b and d),
along with a grayscale intensity plot (c) of the pixels under the yellow double-arrow in d.
The distances in c represent the atomic layer spacing of the unidentified phase from the
(A) 4.5CAC-HH paste sample after 1 year of curing
Three main possible phases were considered for the identity of the Fe-
bearing needle-like phase, described as follows:
i. Al-Fe hydroxide:
After 28 days, Fe signals were detected in Al(OH)3 gels with increasing
proportions of Ca and Si. In literature, amorphous Al(OH)3 were
observed to form polycrystalline "whiskers" morphology shown in the
TEM images in Figure 3.38 a. The different Fe hydroxides (Figure
3.37 a and b; Figure 3.38 b) demonstrated a nano-crystalline needle
morphology, similar to those observed in the 1 year samples. The
mechanism of transformation from the amorphous end-product
ferrihydrite to the needle-like goethite nano-crystals over time, has been
extensively discussed in Bazilevskaya, 2009 [182]. In her study, the
difference in the kinetics of ion dissolution was attributed to the
eventual growth of goethite needles within a poorly-crystalline
ferrihydrite matrix, thereby creating polycrystalline ring diffraction
patterns.
The d-spacing of Al and Fe hydroxides in literature were reported to be
typically below 4 Å [183,184]. However, it has been previously
Figure 3.38 TEM images of (a) Al(OH)3 with its inset selected-area electron diffraction
(SAED) pattern [183]; and of goethite, Fe(OH)O, with its inset SAED pattern [184].
incorporated into the Al-rich hydrate gel, as seen from the TEM and STEM
results. At 28 days, the estimated slag hydration degrees of formulation A
and B were similar (10 % and 9 %, respectively, Figure 3.12) which could
explain the close resemblance of their Mössbauer spectra.
Figure 3.39 57Fe Mössbauer spectrum evolution of formulation A sample over 1 year
of curing. Superimposed are the spectra of formulation B after 28 days of hydration and
of the raw slag.
IS QS MF AA
Sample Component (mm/s) (mm/s) (kOe) (mm/s or kOe) (%)
raw slag (G4) Fe2+ 1.10 2.12 0.00 0.57 39
In Model 1, two new doublets (P1 and P2) were introduced as products of
hydration. The parameters of the P1 component at 7 days, displaying
relatively low IS and QS values, could represent a phase with a similar
structure as the Fe(OH)3 gels reported in PC cement systems [106,135,187]
which in time develops into Fe-monosulfate [129,136]. This
transformation is accompanied by an increased distortion of the atoms
surrounding the Fe signaled by the higher QS values at later ages (<7 days).
However, given the nature (poor to no crystal structure) of the Fe-bearing
hydrate gel observed from TEM and STEM, it was difficult to clearly
postulate the exact atomic environment of the new Fe-containing phases
detected. As Fe is relatively scarce in conventional cement systems,
especially in CAC-based binders, there is no sufficient Mössbauer data on
Fe-containing hydrates, at the moment, to compare with those obtained in
this study. Although the hyperfine parameters of P1 could only be
associated to that of Fe-monosulfate given these constraints, it was clear
from the EDS analyses of TEM and STEM that the Fe-bearing hydrates
could not be monosulfate considering the extremely low (<1.5 at.%)
amount of S detected. Nonetheless, the atomic environment (Fe-Al solid
substitution) and oxidation state (Fe3+) of Fe could be very similar, which
is also the case for the ferrihydrite gels versus the Al-rich hydrate gels at
early ages.
Furthermore, the hyperfine parameters of P2 in Model 1 was even more
intriguing with the combination of IS and QS values that has never been
considered rare [188]. The closest structure that could be compared to this
component is one of the doublets of staurolite
((Fe,Mg)2(Al,Fe)9O6[SiO4]4(O,OH)2) reported in Dyar et al., 1991 [189].
This doublet is in the exceptionally low end of the range for 4-fold
coordinated Fe2+ arising from the delocalization of electrons due to charge
transfer phenomena (i.e., Al3+ Fe2+ or Fe 2+ Fe substitutions) [189]. In
-formed component P2 versus
the Fe2+ in G1 and G2 from the raw slag could suggest a generally more
ordered environment of Fe. This could well be the case if the component
P2 majorly corresponds to the unidentified Fe-bearing phase detected in
STEM, such that the Fe2+ is transitioning from an amorphous state in the
raw slag.
In Model 2, the Fe in the raw slag was assumed to be dominantly Fe2+ (G4)
with an environment comparable to those in fayalite minerals [114,190].
The hyperfine parameters of the doublet P3 varied significantly as the
component shifts from representing a minor Fe3+ proportion of the raw
-bearing hydration product in the paste
samples. This was expected given the simplified approach with minimal
restrictions imposed on the hyperfine parameters and absorption areas, as
described earlier, for Model 2. As hydration proceeds, the best fit obtained
for the spectra suggests P3 parameters closer to those of the Fe-bearing
hydrates in cement systems represented by P1 in Model 1. In addition,
constraint was applied for the different samples, supported the assumption
Figure 3.41 Absorption areas (AA) calculated from the two modelling approaches
By following the changes in the sum of the absorption areas (AA) of the
reactive components of the raw slag (G1, G2 and G3 for Model 1 and G4
for Model 2), the slag hydration degrees of formulation A at different
curing ages were estimated (Figure 3.42). Similar to the assumption
applied in the XCT analysis in section 3.2.2, this estimation equates slag
dissolution to slag hydration. Note that the original equation for the slag
hydration degree presented in Chapter 2 (section 2.2.5) needed correction
due to the apparent overestimation of wüstite especially in Model 2. The
details of these calculations can be referred to in Appendix H.
Figure 3.43 (a) Heat flow and (b) cumulative heat release of pastes containing 23 wt%
paper residue (in blue) and 30 wt% slag (in black).
One concern on the use of this paper residue as SCM is its elevated
(23 wt%) free lime content (Appendix B) excessively surpassing the 1 or
2.519 wt% limit imposed by the EN-450 [191]. Although this limit was
intended for fly ash addition, De Shepper et al., 2011 [192] reiterated the
significance of maintaining the free lime content below 3 wt% in cements
to prevent undesirable effects such as volume expansion, flash setting and
reduced mechanical strength. At 23 wt% addition in the binder, the
effective free lime content was 5.2 wt% from the paper residue alone. For
this reason, this residue was evaluated separately from the other potential
SCM in this study.
In an attempt to increase the slag reactivity in formulation A, the paper
residue was added at 5 wt%, following the blend formulation A-TC (Table
2.3, where TC stands for Top-Crete® paper residue). Even with this small
amount of paper residue, the addition of 0.3 wt% citric acid was needed to
prolong the setting time (Figure 3.44) allowing the freshly-mixed blends
to be properly placed in the moulds. The properties of this blend were then
compared to those of formulation A, and a reference formulation
(reference no TC) with the citric acid addition but without the paper
residue.
Figure 3.44 Initial and final setting time of formulations A-TC; reference no TC; and
formulation A
19
when a complementary autoclave expansion test yields satisfactory results
citric acid addition in improving the workability has been reported [193
195] to have an impact on the mechanical strength, mainly by altering the
microstructure and the porosity during the early age hydration.
Nonetheless, as previously highlighted in Chapter 2, such a low amount of
addition (0.3 wt%) for the binders in this study should not have an
influence in the long term phase assemblage of the binders [82].
Figure 3.45 Compressive strength (primary axis) and SAI evolution (secondary axis) of
formulation A-TC (blue), its reference formulation (gray) and formulation A (black).
SAI was calculated by dividing the strength values of A-TC and the reference by that of
the formulation A.
Figure 3.46 Flexural strength (primary axis) and SAI evolution (secondary axis) of
formulation A-TC (blue), its reference formulation (gray) and formulation A (black).
SAI was calculated by dividing the strength values of A-TC and the reference by that of
the formulation A.
The SAI values of A-TC, particularly those for the compressive strength
(Figure 3.45), suggest that the positive contribution of the paper residue
addition to the strength was most pronounced in the first day of curing,
progressively diminishing over time. This could be related to the length
variation in Figure 3.47 indicating a period of expansion from the 5th hour
Figure 3.47 Length variation of formulations A-TC (blue) and A (black) obtained from
Walter+Bai shrinkage test
Figure 3.48 The slag hydration degree (primary axis) and slag total volume of
formulations A-TC and A estimated from XCT + volume analysis
Figure 3.49 (a) XRD patterns; (b) Mössbauer spectra; (c) DTG curves; and (d) total
bound water evolution of formulations A-TC, its reference and formulation A.
Despite the minimal addition, the early hydration reactions of the paper
residue appeared to have altered the phase assemblage in the first 28 days
by favoring the formation of AFm phases (likely monosulfoaluminates
given the peak locations), evident from the regions marked in red in the
XRD (a) and DTG (c) patterns in Figure 3.49. This is also consistent with
the generally lower total bound water content (d) of formulation A-TC at
1 and 7 days. Similar to the previous observations in formulation A, Fe
was detected only in Al-rich hydrated phase at long curing times (Figure
3.50). After 1 year, the abundance of these hydrate gel phases was apparent
throughout the TEM frames observed.
Figure 3.50 TEM images and the atomic composition of marked areas in the (A-TC)
4.5CAC-HH-5TC paste sample
While the observed increase in strength in the high-sulfate system
(formulation B) was attributed to the increased ettringite formation, that in
A-TC could be partly attributed to a marginally higher degree of slag
hydration based on the XCT estimations (Figure 3.48). These initial
results could boost the interest in the valorization of the paper residue as
an SCM and an activator for slag. In the end, the mix design selection will
be critical in maximizing the synergistic benefits from the slag and the
paper residue.
Conclusion
This chapter demonstrated the reactivity of the slag in binary CAC-HH
systems. The long term (>28 days) contribution of the slag to the
mechanical strength of mortars was confirmed by the evolution of SAI
values increasing over time for both the low (formulation A) and the high
(formulation B) sulfate systems. Although both produced dimensionally-
stable mortars, an overall shrinkage was observed in formulation A as
opposed to the expansive nature of formulation B. This was associated to
the difference in the phase assemblage of these formulations, generally
described using XRD, TGA and ATR-FTIR as follows:
Formulation (A) 4.5CAC-HH: dominated by monosulfoaluminate;
ettringite and gibbsite content decreased over time while strätlingite
formed significantly from 7 days onwards
Formulation (B) 1.6CAC-HH: dominated by ettringite and gibbsite;
with small amounts of strätlingite and monosulfoaluminates
Despite the higher strength values of formulation B mortars, the rate of
slag dissolution remained comparable to that of the low-sulfate system (A),
according to XCT. After 1 year, approximately 49 wt% of the slag
dissolved in formulation A and 43 wt% in B. These estimations were lower
compared to those obtained from Mössbauer spectroscopy. For instance,
56 wt% of the slag was estimated to have reacted after 180 days based on
the latter, while this value was only at 36 % using XCT. This difference
was mainly attributed to the voxel resolution setting (2.15 µm) in XCT,
which could have underestimated the actual values.
No Fe-containing hydrated phase was detected from XRD, TGA and ATR-
FTIR. Using TEM and STEM, Fe was found to be incorporated at low
2-5 at%) in an amorphous phase rich in Al. As the slag
dissolved more significantly after 90 days, an unidentified needle-like
phase with a higher Fe
contained small amounts of Ca, Si and Mg, and could be similar to Al-Fe-
hydroxides or Fe-LDH. Unfortunately, the techniques presented proved to
be insufficient for identifying this phase, given the problems with
intermixing on the nano-scale, and with the sensitivity of hydrates to EDS
analysis. Nonetheless, it was clear that Fe was not detected in the more
conventional hydrated phases, such as ettringite, monosulfoaluminate or
strätlingite.
Moreover, the evolution of the Mössbauer spectra of formulation A
provided strong evidence of the changes in the atomic environment of Fe
during hydration. The 4/5-fold coordinated Fe2+ dominating the raw slag
was transformed into a combination of Fe3+ and Fe2+ in octahedral and 4/5-
Chapter 4
Fe-rich slag addition in ternary CAC-
HH-PC binder
Contents
Introduction 139
Conclusion . . 179
Introduction
In this chapter, the reactivity of the Fe-rich non-ferrous metallurgy (NFM)
slag, selected during the screening test in Chapter 2, was investigated in
ternary calcium aluminate cement (CAC)-calcium sulfate hemihydrate
(HH)-Portland cement (PC) binders. The incorporation of PC in these
CAC-rich systems serves two main purposes:
To emulate a system (CAC-HH-PC) closer to the industrial
formulations of CAC-based blends for rapid repair applications
To evaluate the potential of PC as an activator of the slag
Although there have been several studies on the ternary CAC-HH-PC
binders, detailed in Chapter 1, these preceding investigations were limited
to PC-rich (60-80 wt%) systems. By introducing the slag in formulations
dominated by CAC and HH, the lower carbon footprint and special
properties (i.e., self-drying capacity, rapid hardening) of such binders are
better maximized. Moreover, PC is evaluated as a possible activator of the
slag in these ternary binders. In Chapter 3, the addition of 5 wt% paper
residue slightly increased the rate of slag dissolution while significantly
enhancing the mechanical strength of the mortars. However, due to the
exceptionally elevated free lime content (23 wt%) of this residue, higher
incorporation was not recommended. On the contrary, PC can provide
additional Ca2+ and increase the alkalinity of the hydrated system without
significantly contributing to the free lime content of the raw blend.
The properties of formulations A-PC and B-PC (Table 2.3) were examined
in this chapter. Both of these formulations contained 30 wt% slag and 30
wt% PC. The CAC/HH ratio in formulation A-PC was 4.5 just like in
formulation A. The CAC/HH ratio in formulation B-PC was 1.6 adapted
from formulation B. Several techniques used in Chapter 3 were similarly
employed in this chapter to characterize the hydrated blends over time.
Mechanical strength and dimensional stability were monitored using
mortar samples. The early hydration reactions were compared using
isothermal calorimetry. The degree of slag hydration was estimated using
X-ray Computed Tomography (XCT). The phase assemblage evolution
was followed using XRD, TGA and ATR-FTIR of paste samples. Finally,
Mössbauer spectroscopy, TEM and STEM were used to describe the fate
of Fe during the hydration.
Figure 4.1 Initial and final setting time of formulation A-PC, B-PC and their references
After a high and rapid expansion during the first hour, the momentary
period of shrinkage between the second and third hour in the Walter+Bai
test (Figure 4.4 a) could be an initial indication of ettringite to
monosulfoaluminate conversion as the sulfates were depleted in the
system. This was however simultaneous to other factors affecting the
dimensional changes at early hours such as self-desiccation and drying
shrinkage.
Despite the continuous expansion of the blends during the 2 months of
curing under water (reaching up to 2.3 mm/m for the reference cement),
the mortars appeared to be generally stable and exhibited strength
development. The slag addition seemed to have improved the dimensional
stability of the blend by mitigating the expansion, consistent with its
behavior in the binary systems. With no visible macro cracks signaling
excessive expansion in this low calcium sulfate system (see Figure 4.8),
the SAIcement values were deemed to be acceptable. Nonetheless, the
increasing SAIfiller values confirmed the positive contribution of the slag
addition to the strength, as a reactive, rather than an inert, component.
Less difference in the trends was observed for underwater curing condition
(Figure 4.7 b), where the shape of the curves remained consistent even at
varying expansion degrees. Relative to what was observed from the
reference fillers, it can be said that the stability of the slag-containing
mortars was not exclusively due to binder dilution (less CAC and HH
proportion in the mortars). Instead, the dissolution of the slag likely had an
impact on the phase assemblage and/or the microstructure, manifesting
through the dimensional stability.
Macro cracks started to become apparent after 45 days for the reference
mortars cured under water, where the length variations surpassed
13.7 mm/m (Figure 4.7 b). It is important to mention that despite the
absence of macro cracks and the superior mechanical strength performance
of the slag-containing mortars, the expansion observed in this formulation,
reaching 8 mm/m, was far beyond the standard expansion limits reported
for sulfated PC and CSA systems typically between 0.04 % (0.4 mm/m)
and 0.5 % (5 mm/m), respectively [85,199]. Nonetheless, this degree of
expansion was still below the reported crack threshold limit of 2.25 % [85]
in binary CAC-HH systems, demonstrating the exceptional property of
these CAC-rich blends as expansive binders. Considering that there was
no extensive mix design optimization (i.e., to determine the ideal w/b ratio
and the proportion of each component) and the admixture used was limited
to citric acid, the rapid setting and hardening properties and the expansive
nature of the slag-containing formulations could be considered as
remarkable properties for fast repair applications.
Figure 4.8 Pictures of A-PC and B-PC mortars with their references
after 45 days of curing under water
Figure 4.9 Illustration of the steps involved in the volume analysis of XCT scans
using samples after 28 days of hydration
The plot profiles (b) shown in Figure 4.9 correspond to the 16-bit
grayscale indices of the pixels under the red arrow drawn in their
corresponding 2D reconstructed images (a). The same lower threshold
index limit was used as in the samples from the binary CAC-HH binders
in the previous section. The parameters of the scans and the image analyses
were likewise kept constant for all the samples in this study. It is important
to reiterate that the estimation uses the total volume of the unreacted slag
particles at 1 day as the zero reactivity reference (see section 2.2.4) such
that any slag dissolution occurring before this period were assumed to be
negligible and consequently not accounted in the estimations.
Similar to the scans for formulations A, B and A-TC, high grayscale
contrasts were observed between the slag particles and the rest of the
cement and hydrate matrix even with the addition of PC. It was evident
that the slightly higher density of PC (3.11 g/cm3) compared to that of the
CAC (3.02 g/cm3
(1.4 % Fe2O3 for PC versus 1.8 % Fe2O3 for CAC) and thus yielding a
grayscale index comparable to that of the latter. The combination of these
factors preserved the advantage of the distinct properties of the slag in
using XCT to estimate the degree of slag hydration.
Figure 4.10 presents the estimated degree of slag hydration in
formulations A-PC and B-PC along with those in formulations A, B and
A-TC discussed in the preceding chapter. Note that the estimation was only
performed after 28 days for formulation A-PC and thus the absence of
some data points in the figure.
Figure 4.10 Estimated slag volume and hydration degree using XCT + volume analysis.
Overall, higher degree of slag hydration was estimated for the formulations
incorporating PC as illustrated in Figure 4.10. In the first 28 days, the XCT
results suggest a considerable increase from 9 % (B) to 17 % (B-PC) of the
amount of the dissolved slag with the addition of 30 wt% PC while keeping
the CAC/HH ratio constant at 1.6. Although this increase was less
significant between formulations A and A-PC (4.5 CAC/HH ratio), the
seemingly improved slag reactivity might be considered as a good
direction in improving the 28-day performance of the binders to address
the goal of early-age slag activation. The mechanism could be similar to
what was observed for the paper residue in section 3.5, with the PC
increasing the pH and contributing ions, particularly Ca2+, which
participate in hydrates formation. However, the increased reactivity
became less pronounced beyond 28 days with the differences approaching
50%) towards 1 year.
It is also critical to take into account the possible consequence of the
observed expansion in these PC-containing binders. Although the paste
samples were not subjected to underwater curing, unlike the mortars, the
w/b ratio used was higher (0.65 vs 0.50) for the pastes, failing to exclude
the possibility of expansive behavior. Substantial expansion occurring
after the first day of hydration could result in overestimation of slag
hydration degree as the unreacted slag particles initially located close to
the borders of the assigned scan volume may be slightly displaced out of
the frames. In effect, less unreacted slag volume is estimated not because
of the increased slag reactivity but because of the expansion. In order to
address this possible issue in future analyses of expansive binders, it is
recommended to run a full scan of a smaller sample, rather than setting
fixed dimensions of scan volume which is more ideal for bigger samples
with irregular dimensions.
Nonetheless, the impact of paper residue addition (where shrinkage is
observed) to the estimated slag hydration degree is a clear indication that
the increase in lime content and pH were beneficial in marginally
improving the early-age reactivity of slag according to the XCT
estimations. A similar case is expected with the PC addition despite any
possible contribution of the expansion to the estimated values.
Furthermore, higher slag reactivity was noted in the ternary formulation
with higher sulfate content (B-PC) in the first 28 days of hydration. This is
in contrary with the overall trend observed in the binary systems (A and
B). With these observations, the phase assemblage evolution of both the
systems are discussed in the subsequent section in order to further describe
the possible impact of the blend proportions to the reactivity of the slag.
Figure 4.11 Phase assemblage of the reference (no slag) ternary CAC-HH-PC binder
with varying CAC and HH proportions calculated from thermodynamic modelling
Figure 4.12 Phase assemblage of the ternary binder incorporating 30 wt% slag
and 30 wt% PC with varying CAC and HH proportions calculated from thermodynamic
modelling. Only 50 wt% of the amorphous fraction of the slag was assumed to hydrate.
Equation 4.2
Equation 4.3
where x is equal to 2 for gypsum, 0.5 for hemihydrate and 0 for anhydrite
The models suggest that the addition of slag favors C-S-H formation over
that of ettringite and strätlingite, throughout a wide range of CAC/HH ratio
(Figure 4.12). This is even more evident in the high-sulfate formulation
B-PC for which the model displays a significant proportion of C-S-H along
with monosulfoaluminates (Ms) and strätlingite as dominant phases. In
the A-PC model, on the other hand, the assemblage is dominated by
strätlingite, Ms, and Fe-containing siliceous hydrogarnet. In both
formulations, Fe is incorporated in siliceous hydrogarnet and in an Fe-
containing Ms (Fe-Ms).
- -
4A$H12
1 wt%) Fe-Ms, C4(A,F)$H12. In the latter, the solid solution allows a
variation of the Al/(Al+Fe) ratio between 0 and 0.45. The stability of this
phase is described in detail in Dilnesa et al., 2011 [132]. As this phase was
not suggested to form in the previous models for the binary CAC-HH
systems (Chapter 3), its formation is likely associated to the increased in
Ms (Equation 4.2) and strätlingite (Equation 4.3) proportions resulting
from the instability of CH and gibbsite with the addition PC. In addition,
minor amounts of calcite and O-H hydrotalcite were consistently present
throughout the range to accommodate the C and Mg, respectively, coming
from the PC.
Figure 4.13 Composition by volume of the reference (no slag) ternary binder
with varying CAC and HH proportions simulated from GEMS.
Figure 4.14 Composition by volume of the ternary binder incorporating 30 wt% slag
and 30 wt% PC with varying CAC and HH proportions calculated from GEMS.
Figure 4.15 Heat flow measured ex-situ using isothermal calorimetry for A-PC
(solid black curve) and B-PC (dotted blue curve) pastes
Despite the differences in the heat release curves (Figure 4.15) between
formulations A-PC and B-PC, their cumulative heat release beyond 40
hours remain comparable. This observation is consistent with the trends
observed between the low (A) and high (B) sulfate systems in the binary
blends in Chapter 3.
Figure 4.17 XRD patterns of formulation A-PC (solid black lines) and its reference.
Phases are abbreviated as follows: (Str) strätlingite, (Ett) ettringite, (Ms)
monosulfoaluminate, (Geh) gehlenite, (Cc) calcite, (CSH) C-S-H, (CH) Portlandite,
(Hdg) hydrogarnet and (Wue) wüstite. Note that some of the phases are indicated only to
mark the position of their main peaks, despite their absence in the phase assemblage
Despite the variations mentioned above, there was no drastic change in the
phase assemblage nor was there an appearance of new peaks that could be
directly associated to the hydration products of the slag in A-PC. Only
14 % of the slag has reacted after 28 days according to the XCT results
(Figure 4.10). This is likewise reflected through the stable ATR-FTIR
spectra over time (Figure 4.20) for this formulation. On the contrary, the
changes in formulation B-PC, which contains higher amount of sulfates,
are more apparent particularly between 1 and 7 days. The actual phase
assemblage also appears to be ettringite-rich, in contrary to the
thermodynamically-predicted system rich in monosulfoaluminates. This
could be explained by the lower slag reaction in the real system yielding
an assemblage closer to what was observed for the reference formulation.
Nonetheless, the model was able to predict the instability of CH and
gibbsite in these systems.
Strong ettringite peaks from the XRD patterns (Figure 4.18) alone are
clearly evident, primarily resulting from the fast hydration of CA (from the
CAC) and calcium sulfate (from the HH). The ettringite formed appears to
Figure 4.18 XRD patterns of formulation B-PC (solid blue lines) and its reference.
Phases are abbreviated as follows: (Ett) ettringite, (AFm) AFm, (Ms)
monosulfoaluminate, (Gib) gibbsite, (Geh) gehlenite, (Gyp) gypsum and (Wue) wüstite.
Figure 4.19 DTG curves of (a) formulation A-PC and (b) formulation B-PC
along with their corresponding references. Slag-containing formulations are shown in
solid lines. Phases are abbreviated as follows: (Ett) ettringite, (CSH) C-S-H, (AFm) AFm,
(Str) strätlingite, (Gib) gibbsite, (Hdg) hydrogarnet
Figure 4.20 ATR-FTIR spectra of formulation A-PC (solid black lines) and its reference.
phases identified.
Figure 4.21 ATR-FTIR spectra of formulation B-PC (solid blue lines) and its reference.
phases identified.
Table 4.1 FTIR band assignments based on the distinct spectral peaks observed
Primary Wavenumber
Bond Reference
contributor (cm-1)
O-H Ms/Ett/Str 3432 Horgnies et al., 2013 [137]
1660 Horgnies et al., 2013 [137]
S-O Ett/Ms 1115 Horgnies et al., 2013 [137]
Si-O-Al Str 1150 Okoronkwo and Glasser , 2016 [165]
710 Zapanta et al., 2020 [171]
Having estimated the highest slag hydration degree using XCT for B-PC
as against all the other formulations studied, the ATR-FTIR spectrum of
the raw slag is superimposed to that of formulation B-PC at 1 day in Figure
4.21
that of the B-PC in order to estimate the possible contribution of the
unreacted slag to the peak intensities of the hydrated blends. The
unadjusted spectrum of the raw slag was previously presented in Figure
2.13. Notice that the Si-O asymmetric stretching band of the raw slag at
around 930 cm-1 coincided with the region (between 857 and 965 cm-1)
where the difference between the spectra of the reference and the slag-
containing formulations are observed at 1 day. Assuming that the slag has
barely reacted at this period, it could be assumed that the raw slag is
contributing to the spectrum of the blend in a similar manner to the one
illustrated.
In section 1.4.3, it was highlighted that an evidence of slag reactivity was
depicted in Siakati et al., 2020 [117] principally through peak shift of the
Si-O asymmetric stretching band (700-1200 cm-1) towards a higher
wavenumber. The disappearance of the gaps (red double arrows) in the
superimposed spectra in Figure 4.22 could be interpreted as a possible
consequence of a similar peak displacement towards the higher
wavenumber, causing the unreacted slag peaks to be concealed in the
adjacent areas. Although this could have been accompanied by a shift of
the Si-O rocking band (400-500 cm-1) towards a lower wavenumber, it
could not be clearly distinguished due to the very weak intensity of the
second peak after the dilution in the blend.
Figure 4.22 Superimposed ATR-FTIR spectra of the raw slag (red) to those of the B-PC
at different curing times. It aims to highlight how the gaps formed over time (red double
arrows), could be an evidence of the proposed peak-displacement scenario signifying slag
hydration.
28 days (34 g) and 165 days (47 g) is in agreement with the slag hydration
degree quantified though XCT, discussed earlier in section 4.2. For this
reason, a greater focus is dedicated on this formulation in the subsequent
sections for the investigation on the fate of Fe particularly between 28 to
165 days of curing.
Figure 4.24 Total bound water content of formulations A-PC, B-PC and their references
4.4.1. TEM
The general phase assemblage of formulation A-PC is reflected in the TEM
images shown in Figure 4.25 and Figure 4.26. In these images, the
abundance of strätlingite is apparent with its distinct plate-like morphology
and siliceous composition marked in blue. Most of the particles are
agglomerated in Figure 4.25 a and b, making it difficult for them to be
analyzed individually.
Figure 4.25 TEM images and the atomic composition of marked areas in the (A-PC)
Figure 4.26 TEM images and the atomic composition of marked areas in the (A-PC)
4.5CAC-HH-30PC paste sample after 28 days of hydration
Figure 4.27 TEM images, diffraction pattern and the atomic composition of Fe-
containing gels in the (B-PC) 1.6CAC-HH-30PC paste sample after 28 days of hydration
Figure 4.28 TEM images and the atomic composition of marked areas in the (B-PC)
1.6CAC-HH-30PC paste sample after 150 days of hydration
The Fe content after 150 days of curing was generally elevated as the
hydrated gels (B-PC 8-11) appear to be thicker in Figure 4.28. This was
accompanied by an increase in the atomic percentages of Ca and Si almost
twice as those obtained for the gels at 28 days. It is probable that the late
dissolution of PC and slag contributed to this increase. Since CAC was
reacting immediately in the early days of curing, the Al content of the gels
remained stable between 28 and 150 days.
Once again, no significant amount of Fe was detected in ettringite (B-PC
14) even after longer curing time. Unreacted wüstite was detected at 150
days (Figure 4.28 B-PC 13) exhibiting its distinct dendritic morphology.
This supports the earlier remark that only the amorphous fraction of the
slag is dissolving during the hydration. Adjacent to the Fe-rich wüstite (35
at% Fe in Figure 4.28 e), a hydrated gel (B-
distinguished with a highly contrasting Fe signal. This demonstrates the
precision of the EDS in analyzing different phases on a nano-scale.
Although the intermixing of phases due to the complex nature of the
4.4.2. STEM
In Chapter 3, the elemental maps obtained from STEM demonstrated the
non-uniform distribution of Fe in the hydrated gels. It has driven the
proposition that there are at least two Fe-containing hydrate phases in the
CAC-based systems: (1) an amorphous part with a lower Fe concentration;
(2) and an unidentified nano-crystalline phase comprising the darker
regions of the gels, with a higher Fe content. Although the TEM results in
the preceding section indicated a different composition of the gels in the
ternary binders, the inhomogeneity of these gels provided an initial
evidence of the possible presence of intermixed phases congruent to the
binary systems.
Figure 4.29 STEM elemental mapping of a hydrated gel phase in the (B-PC)
1.6CAC-HH-30PC paste sample after 150 days of curing. In analyzing the STEM
elemental maps, note that the color intensities were independent for each element.
Figure 4.30 STEM elemental mapping for a selected area in the (B-PC)
1.6CAC-HH-30PC paste sample after 150 days of curing.
Figure 4.31 STEM images and their corresponding Fe and S maps from (B-PC)
1.6CAC- HH-30PC paste sample after 150 days of curing.
Figure 4.32 Ternary diagram of Fe-containing hydrated gels with 1-12 at% Fe observed
using TEM and STEM. The values were calculated from the Ca, Si and Al atomic
composition from EDS analyses normalized to 100 at%. Reference points for, gibbsite,
strätlingite, katoite and Si-hydrogarnet were added in the diagram based on their
empirical formula while the C-A-S-H region was estimated based on the reported stable
composition in Portland cement system [202]. The area covered by the Al-rich gels in
component (M2) was used only for the raw slag fitting. The modeled
spectrum of the B-PC paste after 28 days is shown in Figure 4.34.
Figure 4.33 57Fe Mössbauer spectrum evolution of formulation B-PC after 28 days
along with the spectra of the raw slag and the raw PC used in the formulations. Running
average per 10 data points was used to fit a curve for the PC spectrum (section 2.2.5) in
order to minimize the noisy signature resulting from its low (1.4% Fe2O3) Fe content.
Figure 4.34 57Fe Mössbauer spectrum of formulation B-PC at 28 days and its
deconvoluted components
The absorption areas (AA) obtained from the fitted models at different
curing time are presented in Figure 4.35. Using the same equation
elaborated in Appendix J, the degree of slag hydration was estimated based
on the changes in the sum of the AA of the unreacted slag components.
The results of this estimation is presented in Figure 4.36.
Figure 4.35 Evolution of B-PC absorption areas (AA) calculated based on the models.
Values at 0 day corresponds to the fitting of the raw slag.
The results confirm the earlier remark on the resemblance of the B-PC
spectrum at 28 days (with 42 % slag hydration degree) to that of the A
spectrum at 90 days (with 48 % slag hydration degree). The slower
increase in the values between 1 and 28 days relative to the trend for the
binary systems in section 3.4.3 suggests that a similar extent of dissolution
was achieved faster in the ternary blends with PC. This is evident from the
steep slope between 0 and 1 day in Figure 4.36.
Figure 4.36 Slag hydration degree of formulation B-PC (primary axis) estimated based
on the changes in the total absorption areas (secondary axis) of the unreacted slag
Nonetheless, it is important to remark the huge difference in the slag
hydration degree values obtained from this estimation compared to those
from the XCT in section 4.2. As mentioned for binary systems,
underestimation from XCT could be linked to two main factors: (1) the
assumption that the slag is completely unreactive in the first 24 hours of
hydration; and (2) the 2.15 µm voxel resolution setting which excluded the
contribution of the finer slag particles. This could be addressed in the
future works by performing faster scans immediately after the paste
hardens and several others within the first 24 hours of curing. Sensitivity
analysis on the voxel resolution setting could also be performed in order to
quantify the influence of this factor to the estimation.
Finally, the components of the fitting procedure followed for the
Mössbauer spectroscopy could be further refined and adapted for the
ternary systems. Although Fe was detected in intermixed hydrated gels in
both the binary and ternary systems, the difference in their compositions
merits a re-evaluation of the fitted components, P1 and P2. Further
characterization preferably using synthetic systems with less components
will be a critical prerequisite in identifying the better adapted components
for the re-evaluation.
Conclusion
This chapter demonstrated the reactivity of the slag in ternary CAC-HH-
PC systems. The binders exhibited an expansive nature, particularly those
of formulation B-PC which had the higher sulfate content. In comparison
to the reference mortars, the addition of the slag proved to be beneficial in
mitigating excessive expansion such that macro cracks and consequently
the decrease in mechanical strength were avoided even when the mortars
were cured under water. In addition, promising SAI values (>1) exceeding
those obtained in the binary system were achieved. High compressive
strengths at around 20 MPa were obtained at 1 day, which increased to
more than 40 MPa at 28 days for the slag-containing mortars. However,
the addition of the PC significantly accelerated the setting time prompting
the need for 0.3 wt% citric acid addition to provide sufficient time for the
mixtures to be placed in the moulds. This resulted in setting time ranging
from 20-40 min.
Estimation of the degree of slag hydration using Mössbauer spectroscopy
and XCT both supported the faster kinetics of slag hydration with the
addition of PC. The reactivity of the slag in the first 28 days of hydration
was reflected through the drastic changes in the Mössbauer spectra,
particularly of the high-sulfate formulation. On the other hand, the XCT
results showed that although the slag was more reactive for the ternary
binders in the first 28 days, the difference in the hydration degree was less
pronounced at a longer curing time. Despite the consistency in the trends
observed between these two techniques, however, significant differences
in the estimated values were linked to two main factors: (1) the assumption
that the slag is completely unreactive in the first 24 hours of hydration; and
(2) the 2.15 µm voxel resolution setting which excluded the contribution
of the finer particles.
The phase assemblages obtained experimentally based on the XRD, TGA
and ATR-FTIR results could be summarized as follows:
Physical Properties
As presented in Figure 5.1, high compressive strengths were obtained for
mortars incorporating the slag and 1.3 wt% citric acid. The latter, initially
added to delay the setting, also assisted in the early strength development.
Overall, higher compressive strengths were achieved in the high-sulfate
systems due to the more extensive ettringite formation. However, the
expansive nature of ettringite proved to be problematic in reference cement
mortars containing PC (formulation B-PC) for which excessive expansion
eventually led to macro cracks (Figure 5.2) and loss in strength. The
addition of slag mitigated this expansion, producing more dimensionally-
stable mortars. This is likely governed by the dilution effect (physical
factor) in the early hydration ages, and by the resulting phase assemblage
(chemical contribution of the slag) at long term.
Figure 5.2 Pictures of the B-PC and B (high-sulfate) mortars after 45 days of curing
under water
Figure 5.3 Consolidated slag hydration degree estimated using XCT and Mössbauer
spectroscopy
Both results supported the faster kinetics of slag hydration with the
addition of PC. At longer curing period however, the estimated values all
converged towards 50 %. The generally lower values obtained from XCT
were linked to two main factors: (1) the assumption made in XCT that the
slag is completely unreactive in the first 24 hours of hydration; and (2) the
2.15 µm voxel resolution setting in XCT which excluded the contribution
of the finer slag particles or of other phases present at this scale.
While the experimental results are generally in agreement with the models,
the differences were attributed predominantly to the kinetics of hydration,
not accounted for during the simulation. Experimentally, ettringite
dominated the high-sulfate systems in the early hydration ages, while the
slag and PC exhibited slower hydration kinetics relative to the CAC and
HH. The ettringite formed was stable over time, therefore modifying the
proportion of the ions available for hydration at later ages. This could also
be the reason why the Fe-containing hydrogarnet and Fe-
monosulfoaluminate predicted in the models were not detected
experimentally.
Following the experimental phase assemblage evolution over time, the
influence of the slag was manifested in the silicates (i.e., strätlingite,
gehlenite, C-S-H) regions where the major changes were observed, relative
to the reference formulations. The crystalline component of the slag,
wüstite, also appeared unreactive over time, demonstrating that only the
glass fraction was dissolving throughout the 1-year-curing period.
Moreover, another evidence of slag reactivity was proposed using the
ATR-FTIR spectra through a hypothetical peak-displacement of the raw
-O asymmetric stretching band at 930 cm-1. However, the
complexity of the systems studied and the difficulty in characterizing the
amorphous phases did not allow a conclusive answer as to the fate of Fe
using the more common characterization techniques in cement hydration.
Figure 5.4 Consolidated ternary diagram of Fe-containing hydrate gels with 1-12% Fe
obtained from (TEM-) EDS analyses. Reference points for, gibbsite, strätlingite, katoite,
Si-hydrogarnet and C-A-S-H were added from literature data [202].
Future Outlook
The extensive characterization of the physical and mineralogical properties
of the CAC-based binders incorporating some of the SOCRATES residues,
demonstrated the positive potential of this valorization pathway.
Nonetheless, better understanding of the hydration mechanism of the slag
will be needed to eventually adapt the properties of the binders to their
target applications.
Future works should focus on complementary analyses to identify the Fe-
bearing nano-crystalline phase detected using STEM. This could be done
by employing other advanced characterization methods such as nuclear
magnetic resonance (NMR) or X-ray absorption (XAS) spectroscopy,
which have been previously used [113,123,206] in studying the uptake of
Fe in hydrated phases in cement systems. The identification could also be
facilitated by reproducing the phase synthetically and at higher
concentrations, in order to minimize the problem of intermixing on a nano-
scale.
Moreover, the methodology developed for quantifying the degree of slag
hydration using XCT could be optimized by reducing the voxel resolution
and including the finer fractions in the estimation. Although this would
imply a scan with longer duration or smaller volume. A refined selective
dissolution procedure based on salicylic acid-methanol [141] or NaOH-
EDTA [139,207] could also be developed to validate the estimations.
The models used for the fitting of the Mössbauer spectra could be
improved by finding the compromise between the simplistic and complex
models in order to reflect the reality more accurately. The selection of the
fitting parameters could be supported by complementary techniques, such
as titration using H2SO4, HF, and H3BO3 [208] to confirm the oxidation
state of Fe before and after hydration.
Optimizing the proportions of CAC, HH and PC aided by design of
experiment (DOE) could also potentially improve the reactivity of the slag
and the overall properties of the blend. From an industrial point of view,
this could be beneficial as the applicative performance, along with
economic factors, are crucial considerations. Finally, the long-term
durability of the optimized formulation in different chemical environments
must be tested to better understand the behavior of the blend and adapt it
according to the application.
Appendices
A. Life cycle assessment and exergy analysis on the use of industrial
by-products as SCM in PC system
The paper attached in this section was presented by the author at the Forum
of Young Researchers in Sustainable Building 2019 (YRSB19) in Prague,
Czech Republic on July 2, 2019. This work focused on the addition of
fumed fayalite slag (Koranel®), one of the SOCRATES residues, in a pure
PC system. These results were excluded from the main part of the thesis
which involved CAC-rich binders.
Figure B.1 XRD pattern and Rietveld analysis of the granulated Fe-rich slag abbreviated
Figure B.2 XRD pattern and Rietveld analysis of the slowly-cooled Fe-rich slag
-
Figure B.3 XRD pattern and Rietveld analysis of the fumed fayalite slag (Koranel®)
Figure B.4 XRD pattern and Rietveld analysis of the bottom ash
jarosite residue
Figure B.5 XRD pattern and Rietveld analysis of the jarosite residue
Figure B.6 XRD pattern and Rietveld analysis of the paper residue
Cements
CAC, Secar 51®
Figure B.7 XRD pattern and Rietveld analysis of the CAC (Secar 51®) abbreviated as
Figure B.8 XRD pattern and Rietveld analysis of the CAC (Ciment Fondue®)
abbreviated
Figure B.9 XRD pattern and Rietveld analysis of the ordinary Portland cement
Figure B.10 XRD pattern and Rietveld analysis of the calcium sulfate hemihydrate
Figure B.11 XRD pattern and Rietveld analysis of the quartz filler abbreviated
in this study
Figure D.2 Comparison of the flexural strengths of air-cured versus water-cured mortar
samples. The dotted diagonal line represents values at 1:1 ratio
Figure F.1 Illustration of IS, QS and MF using the Mössbauer spectra of wüstite (a) and
magnetite (b)
Total Gaussian-
Statistical parameter related to the QS derived from the least square
minimization fitting in the IMSG software. Gaussian-type symmetric
spreading of the QS values was allowed during the fitting.
Total area (A)
The total area fitted to each Fe component proportional to the quantity
(percent composition) that they represent in the modelled spectrum.
During the fitting, the typical errors for the IS and QS were ± 0.02 mm/s
and ± 5 % for the A. The components selected were based on the results
from complementary characterization techniques and literature data on
synthetic systems. Chi-squared test was used to evaluate the best fit for
each dataset evaluated.
Some minor (< 1 wt%) components of the raw materials (i.e., ZnO,
CuO, P2O5 from the slag) were considered unreactive during the
The components
used could be referred to in the codes provided.
Wüstite from the slag was considered unreactive and was added in the
was allowed to react roughly guided by the estimations from XCT after
1 year of hydration.
Fe-bearing hydrates that were permitted in the equilibrium state include
CSH, ettringite, monosulfates, hydrogarnet, monocarbonate,
ferrihydrites and other AFm and AFt phases. Phases that were
deactivated included pyrite, goethite, elemental iron, hematite,
magnetite, etc.
The mineralogical composition (CAC, HH and PC) and chemical assay
(slag) of each component presented in Chapter 2 were entered in the
modC array columns.
For some models, extra O2 component was added to aid the balance
In the Sampling section of the Process Simulators, the following code was
used to calculate the phase assemblage composition in weight percentages:
xp[J] =: J;
yp[J][0] =: phM[{CSHQ}];
yp[J][1] =: phM[{straetlingite}];
yp[J][2] =: phM[{ettringite}];
yp[J][3] =: phM[{SO4_OH_AFm}];
yp[J][4] =: phM[{OH_SO4_AFm}];
yp[J][5] =: phM[{SO4_CO3_AFt}];
yp[J][6] =: phM[{CO3_SO4_AFt}];
yp[J][7] =: phM[{Gibbsite}];
yp[J][8] =: phM[{C3AH6}];
yp[J][9] =: phM[{CAH10}];
yp[J][10] =: phM[{C4AH11}];
yp[J][11] =: phM[{C4AsH12}];
yp[J][12] =: phM[{Calcite}];
yp[J][13] =: phM[{C3FS0.84H4.32}];
yp[J][14] =: phM[{C3FS1.34H3.32}];
yp[J][15] =: phM[{Gypsum}];
yp[J][16] =: phM[{hemihydrate}];
yp[J][17] =: phM[{Ferrihydrite-am}];
yp[J][18] =: phM[{OH-hydrotalcite}];
yp[J][19] =: phM[{aq_gen}];
yp[J][20] =: 0.01* (modC[10][0]*cNu+ modC[10][1]*(100-cNu));
CAC-HH-slag system
In the Process Simulators of GEMS, the code used for the Controls section
was as follows:
$modC0 S51 modC1 HH modC2 slag
xa_[{Al2O3}] =: 0.01*(cNu*modC[0][0]+modC[0][1]*(70-cNu)+modC[0][2]*30);
xa_[{C2S}] =: 0.01*(cNu*modC[1][0]+modC[1][1]*(70-cNu)+modC[1][2]*30);
xa_[{CA}] =: 0.01*(cNu*modC[2][0]+modC[2][1]*(70-cNu)+modC[2][2]*30);
xa_[{CaCO3}] =: 0.01*(cNu*modC[3][0]+modC[3][1]*(70-cNu)+modC[3][2]*30);
xa_[{CaO}] =: 0.01*(cNu*modC[4][0]+modC[4][1]*(70-cNu)+modC[4][2]*30);
xa_[{CaSO4_05H2O}] =: 0.01*(cNu*modC[5][0]+modC[5][1]*(70-
cNu)+modC[5][2]*30);
xa_[{SiO2}] =: 0.01*(cNu*modC[6][0]+modC[6][1]*(70-cNu)+modC[6][2]*30);
xa_[{C4AF}] =: 0.01*(cNu*modC[7][0]+modC[7][1]*(70-cNu)+modC[7][2]*30);
xa_[{FeO}] =: 0.01*(cNu*modC[8][0]+modC[8][1]*(70-cNu)+modC[8][2]*30);
xa_[{MgO}] =: 0.01*(cNu*modC[9][0]+modC[9][1]*(70-cNu)+modC[9][2]*30);
xa_[{Na2O}] =: 0.01*(cNu*modC[10][0]+modC[10][1]*(70-cNu)+modC[10][2]*30);
xa_[{K2O}] =: 0.01*(cNu*modC[11][0]+modC[11][1]*(70-cNu)+modC[11][2]*30);
xa_[{Aqua}] =: 65;
xa_[{O2}] =: 10;
xa_[{Gypsum}] =:0.01*(cNu*modC[0][0]+modC[0][1]*(70-cNu)+modC[0][2]*30);
xa_[{C3A}] =: 0.01*(cNu*modC[1][0]+modC[1][1]*(70-cNu)+modC[1][2]*30);
xa_[{C3S}] =: 0.01*(cNu*modC[2][0]+modC[2][1]*(70-cNu)+modC[2][2]*30);
xa_[{Al2O3}] =: 0.01*(cNu*modC[3][0]+modC[3][1]*(70-cNu)+modC[3][2]*30);
xa_[{C2S}] =: 0.01*(cNu*modC[4][0]+modC[4][1]*(70-cNu)+modC[4][2]*30);
xa_[{CA}] =: 0.01*(cNu*modC[5][0]+modC[5][1]*(70-cNu)+modC[5][2]*30);
xa_[{CaCO3}] =: 0.01*(cNu*modC[6][0]+modC[6][1]*(70-cNu)+modC[6][2]*30);
xa_[{CaO}] =: 0.01*(cNu*modC[7][0]+modC[7][1]*(70-cNu)+modC[7][2]*30);
xa_[{CaSO4_05H2O}] =: 0.01*(cNu*modC[8][0]+modC[8][1]*(70-
cNu)+modC[8][2]*30);
xa_[{SiO2}] =: 0.01*(cNu*modC[9][0]+modC[9][1]*(70-cNu)+modC[9][2]*30);
xa_[{C4AF}] =: 0.01*(cNu*modC[10][0]+modC[10][1]*(70-cNu)+modC[10][2]*30);
xa_[{K2O}] =: 0.01*(cNu*modC[11][0]+modC[11][1]*(70-cNu)+modC[11][2]*30);
xa_[{Aqua}] =: 65;
In the Process Simulators of GEMS, the code used for the Controls section
was as follows:
xp[J] =: J;
yp[J][0] =: phM[{Gypsum}];
yp[J][1] =: phM[{ettringite}];
yp[J][2] =: phM[{SO4_OH_AFm}];
yp[J][3] =: phM[{straetlingite}];
yp[J][4] =: phM[{Gibbsite}];
yp[J][5] =: phM[{CSHQ}];
yp[J][6] =: phM[{C3AH6}];
yp[J][7] =: phM[{Portlandite}];
yp[J][8] =: phM[{C3(AF)S0.84H}];
yp[J][9] =: phM[{Ferrihydrite-am}];
yp[J][10] =: phM[{Ferrihydrite-mc}];
yp[J][11] =: 0.01*(cNu*modC[12][0]+modC[12][1]*(70-cNu)+modC[12][2]*30);
yp[J][12] =: phM[{aq_gen}];
CAC-HH-PC-slag system
In the Process Simulators of GEMS, the code used for the Controls section
was as follows:
$modC0 Secar51 modC1 HH modC2PC mod3slag
xa_[{Gypsum}] =:0.01*(cNu*modC[0][0]+modC[0][1]*(40-cNu)+modC[0][2]*30+
modC[0][3]*30);
xa_[{C3A}] =: 0.01*(cNu*modC[1][0]+modC[1][1]*(40-cNu)+modC[1][2]*30+
modC[1][3]*30);
xa_[{C3S}] =: 0.01*(cNu*modC[2][0]+modC[2][1]*(40-cNu)+modC[2][2]*30+
modC[2][3]*30);
xa_[{Al2O3}] =: 0.01*(cNu*modC[3][0]+modC[3][1]*(40-cNu)+modC[3][2]*30+
modC[3][3]*30);
xa_[{C2S}] =: 0.01*(cNu*modC[4][0]+modC[4][1]*(40-cNu)+modC[4][2]*30+
modC[4][3]*30);
xa_[{CA}] =: 0.01*(cNu*modC[5][0]+modC[5][1]*(40-cNu)+modC[5][2]*30+
modC[5][3]*30);
xa_[{CaCO3}] =: 0.01*(cNu*modC[6][0]+modC[6][1]*(40-cNu)+modC[6][2]*30+
modC[6][3]*30);
xa_[{CaO}] =: 0.01*(cNu*modC[7][0]+modC[7][1]*(40-cNu)+modC[7][2]*30+
modC[7][3]*30);
xa_[{CaSO4_05H2O}] =: 0.01*(cNu*modC[8][0]+modC[8][1]*(40-cNu)+
modC[8][2]*30+modC[8][3]*30);
xa_[{SiO2}] =: 0.01*(cNu*modC[9][0]+modC[9][1]*(40-cNu)+modC[9][2]*30+
modC[9][3]*30);
xa_[{C4AF}] =: 0.01*(cNu*modC[10][0]+modC[10][1]*(40-cNu)+modC[10][2]*30+
modC[10][3]*30);
xa_[{FeO}] =: 0.01*(cNu*modC[11][0]+modC[11][1]*(40-
cNu)+modC[11][2]*30+modC[11][3]*30);
xa_[{MgO}] =: 0.01*(cNu*modC[12][0]+modC[12][1]*(40-
cNu)+modC[12][2]*30+modC[12][3]*30);
xa_[{Na2O}] =: 0.01*(cNu*modC[13][0]+modC[13][1]*(40-
cNu)+modC[13][2]*30+modC[13][3]*30);
xa_[{K2O}] =: 0.01*(cNu*modC[14][0]+modC[14][1]*(40-
cNu)+modC[14][2]*30+modC[14][3]*30);
xa_[{Aqua}] =: 65;
xa_[{O2}] =: 25;
In the Sampling section of the Process Simulators, the following code was
used to calculate the phase assemblage composition in weight percentages:
xp[J] =: J;
yp[J][0] =: phM[{Gypsum}];
yp[J][1] =: phM[{ettringite}];
yp[J][2] =: phM[{ettringite-AlFe}];
yp[J][3] =: phM[{SO4_OH_AFm}];
yp[J][4] =: phM[{monosulph-AlFe}];
yp[J][5] =: phM[{straetlingite}];
yp[J][6] =: phM[{Gibbsite}];
yp[J][7] =: phM[{CSHQ}];
yp[J][8] =: phM[{CNASH}];
yp[J][9] =: phM[{C3AH6}];
yp[J][10] =: phM[{Portlandite}];
yp[J][11] =: phM[{C3(AF)S0.84H}];
yp[J][12] =: phM[{Ferrihydrite-am}];
yp[J][13] =: phM[{Calcite}];
yp[J][14] =: phM[{C3FS0.84H4.32}];
yp[J][15] =: phM[{C3FS1.34H3.32}];
yp[J][16] =: phM[{MgAl-OH-LDH}];
yp[J][17] =: phM[{OH-hydrotalcite}];
yp[J][18]] =: 0.01*(cNu*modC[15][0]+modC[15][1]*(40-cNu)+modC[15][2]*30+
modC[15][3]*30);
yp[J][19] =: phM[{aq_gen}];
Figure J.2
mponents (secondary axis)
Or simply:
Figure J.6 Mössbauer spectrum and components of the raw slag derived using Model 2
Figure J.11 Mössbauer spectrum and components of formulation A-TC (after 28 days of
curing) derived using Model 2
Figure K.1 XRD patterns of pure PC paste after 1, 7 and 28 days of hydration
From both the XRD and DTG patterns alone, it is clear that the hydrated
system is dominated by portlandite (CH) and the amorphous C-S-H (CSH).
The sharp diffraction peak of the CH easily masked the peaks of the minor
hydrate phases such as calcite, ettringite and other calcium aluminate
hydrate phases. The DTG curves were critical in characterizing the
presence of amorphous C-S-H and in delineating the position of its peak
relative to ettringite: C-S-H at 90-140°C, and ettringite at around 140°C.
The latter proves to be essential in interpreting the patterns of formulations
A-PC and B-PC in chapter 4. In addition, the peaks at around 172°C could
be assigned to the formation of minor amounts of monosulfoaluminates. A
small shoulder at 92°C could be due to residual water or isopropanol
during the sample preparation.
The ATR-FTIR spectra are dominated by the strong Si-O peak at 950 cm-
1
assigned to the asymmetric stretching vibration band of C-S-H. Out of
plane and in plane vibration bands between 450 and 500 cm-1 were also
assigned to C-S-H. The sharp peaks at around 3600 cm-1 are the signature
of the O-H band in portlandite. The diminishing peak at around 1110cm-1
Figure K.2 DTG curves of pure PC paste after 1, 7 and 28 days of hydration
Figure K.3 ATR-FTIR spectra of pure PC paste after 1, 7 and 28 days of hydration
List of Publications
Part of the results of this work has been previously presented by the author
in the following publications:
Journal articles
J. Astoveza; R. Trauchessec; R. Soth ; Y. Pontikes, Properties of
Calcium Aluminate Blended Cement incorporating Fe-rich Slag:
Evolution over a Curing Period of 1 Year . Journal of Construction and
Building Materials. (2021).
https://doi.org/10.1016/j.conbuildmat.2021.122569
J. Astoveza; R. Trauchessec; S. Migot-Choux; R. Soth; Y.
Pontikes, Fe-rich slag addition in ternary Portland cement, aluminate
cement and calcium sulfate binders. Journal of Cement and Concrete
Research (2022).
https://doi.org/10.1016/j.cemconres.2021.106689
Conferences
6th International Slag Valorisation Symposium. Mechelen
(Belgium). 1-5 April 2019 poster presentation and conference paper
entitled: Assessing the reactivity of industrial by-products in calcium
aluminate cement-based formulations
Forum of Young Researchers in Sustainable Building 2019.
Prague (Czech Republic). 2-4 July 2019 oral presentation and
conference paper entitled: Industrial By-Products as Non-Conventional
Supplementary Cementitious Material.
International Process Metallurgy Symposium. Espoo (Finland). 5-
6 November 2019 oral presentation entitled: Quantifying the Degree
of Fe-rich Slag Hydration in Calcium Aluminate Blended Cement by
Image Analysis of SEM-BSE and XCT.
Concrete Solutions towards Carbon Neutral Construction by 2050,
Advanced Materials for Sustainable Infrastructure Development,
Gordon Research Conference. Ventura (USA), 22-23 February 2020
poster presentation entitled: Calcium Aluminate Blended Cement
incorporating Fe-rich Slag: Overview and Characterization
Techniques.
Seminars
6 June
2019 - poster presentation entitled: Calcium Aluminate Blended
Cements Incorporating Engineered Residues.
The submission of a third journal article is planned for the third quarter of
2022. The paper will cover the results of the STEM and Mössbauer
spectroscopy on formulation (A) 4.5CAC-HH.
References
[1] European Commission, Circular Economy Action Plan: The
European Green Deal, (2021), https://ec.europa.eu/.
[2] SOCRATES EU MSCA-ETN, Project proposal: Innovative training
networks (ITN) call: H2020-MSCA-ITN-2016, Part B, (2016),
https://etn-socrates.eu/.
[3] United Nations Environement Program (UNEP), Buildings and
climate change: Summary for decision-makers, (2009),
www.unep.fr/scp/sun.
[4] K.L. Scrivener, V.M. John, E.M. Gartner, Eco-efficient cements:
Potential economically viable solutions for a low-CO2 cement-
based materials industry, Cement and Concrete Research 114 (2018)
2 26.
[5] P.J.M. Monteiro, S.A. Miller, A. Horvath, Towards sustainable
concrete, Nature materials 16 (2017) 698 699.
[6] A. Vollpracht, B. Lothenbach, R. Snellings, J. Haufe, The pore
solution of blended cements: a review, Mater Struct 49 (2016)
3341 3367.
[7] S. Samad, A. Shah, M.C. Limbachiya, Strength development
characteristics of concrete produced with blended cement using
ground granulated blast furnace slag (GGBS) under various curing
1213.
[8] B. Lothenbach, K. Scrivener, R.D. Hooton, Supplementary
cementitious materials, Cement and Concrete Research 41 (2011)
1244 1256.
[9] J. Bizzozero, Hydration and dimensional stability of calcium
aluminate cement based systems, (Doctorate thesis): Faculté des
sciences et techniques de l'ingénieur, Laboratoire des matériaux de
constrcution, Lausanne, EPFL, (2014).
[10] International Energy Agency, Technology Roadmap: Low-Carbon
Transition in the Cement Industry, (2018),
https://iea.blob.core.windows.net/.
[11] J.E. Rossen, Composition and morphology of C-A-S-H in pastes of
alite and cement blended with supplementary cementitious
materials, Lausanne, EPFL, (2014).