Sei sulla pagina 1di 165

Interdisciplinary Nanoscience Center, 2011

FACULTY OF SCIENCE
AARHUS UNIVERSITY
F
l
u
o
r
i
d
e

M
i
n
e
r
a
l
i
z
a
t
i
o
n

o
f

P
o
r
t
l
a
n
d

c
e
m
e
n
t
P
h
D

T
h
e
s
i
s


2
0
1
1
T
h
u
a
n

T
.

T
r
a
n
Front cover illustration. Stacked plot of a 2D
29
Si{
19
F} HETCOR NMR spectrum for
a synthetic sample of cuspidine (Ca
4
Si
2
O
7
F
2
).
PhD Thesis
THUAN T. TRAN
Fluoride Mineralization
of Portland cement
Applications of Double-Resonance NMR Spectroscopy
in Structural Investigations of Guest Ions in Cement Phases












Thuan Thai Tran

PhD Thesis
Instrument Centre for Solid-State NMR Spectroscopy
Interdisciplinary Nanoscience Center (iNANO)
Department of Chemistry, Faculty of Science and Technology
Aarhus University, Denmark

August 2011


Preface

This thesis presents the results obtained during my PhD study from August 2007 to July
2011 at the Instrument Centre for Solid-State NMR Spectroscopy, Department of Chemistry, and
the Interdisciplinary Nanoscience Center (iNANO) at Aarhus University. The PhD study has
been a part of the FUTURECEM project (2007 2010), which is a joint collaboration between
Aalborg Portland A/S, iNANO (Aarhus and Aalborg Universities) and the Geological Survey of
Denmark and Greenland (GEUS), Copenhagen, with partly financial support from the Danish
National Advanced Technology Foundation (Hjteknologifonden).



Acknowledgments

It is a pleasure for me to express my gratitude to all the people
who have contributed with valuable assistances to my PhD project.

First and foremost, I would like to thank my supervisor Associate Professor Jrgen
Skibsted for his excellent support and scientific guidance, ever since I joined the Instrument
Centre for Solid-State NMR Spectroscopy as a young, undergraduate student for almost eight
years ago. He introduced me to solid-state NMR as well as cement research and gave me the
opportunity to extend my studies at the Instrument Centre. His ongoing fruitful discussions on
both subjects have been a great source of inspiration to this PhD project.
Professor Hans Jrgen Jakobsen and Associate Professor Henrik Bildse are
acknowledged for their general interest in the project and several fruitful discussions about
experimental as well as theoretical solid-state NMR problems. They are also thanked for being
available when there have been problems with the spectrometers or the computers. Mrs. Rigmor
Seberg Johansen and Mrs. Anne Birgitte Bundgaard Johannsen are thanked for their valuable
assistance in conducting numerous of NMR experiments during my time at the Instrument
Center. I also want to acknowledge all the members at the NMR laboratory, including PhD Sren
Lundsted Poulsen and Phd students Sren Srensen, Tine Fly Sevelsted, Maiken Rabl
Jrgensen and Nicklas Bromose Kolman, for creating a good scientific and social atmosphere.
I would like to thank the researchers at the R&D, Quality and Technical Sales Support at
Aalborg Portland A/S: Project Managers Lise Frank Kirkegaard and Mette Steenberg Gly and
Chief Scientist Duncan Herfort for a good collaboration and several interesting discussions on
the cement chemistry.
Finally, the Danish National Advanced Technology Foundation and the Faculty of
Science and Technology, Aarhus University are acknowledged for their financial support of this
PhD project.



Abstract
The increasing applications of alternative raw materials, alternative fuels and
supplementary cementitious materials (SCMs) in todays cement production lead to a significant
amount of impurities (e.g. fluoride, SO
3
, P
2
O
5
etc.) in the final cement materials. These so-called
minor components may also be added to the raw meal of starting materials to modify the
clinkering processes. They introduce several important effects on the formation of the cement
clinker phases as well as on their hydrational properties. In general, the structural
characterization of the minor components is rather difficult due to their low quantities in cement.
This PhD thesis focuses on two main subjects: (i) modification of the Portland clinker
composition with the aim of increasing the clinker reactivity, enabling a 30 % replacement of the
Portland clinkers by SCMs without a significant reduction in the strength properties of the
blended cement, and (ii) development and implementation of solid-state double-resonance
Nuclear Magnetic Resonance (NMR) techniques including Cross Polarization (CP), Rotational-
Echo Double-Resonance (REDOR) and Rotational-Echo Adiabatic-Passage Double-Resonance
(REAPDOR) experiments in the structural characterization of cementitious materials.
The mineralizing properties of calcium fluoride have been investigated in detail. The
mechanism for fluoride mineralization is probed using different solid-state
19
F,
27
Al and
29
Si
NMR techniques. Furthermore, the impacts of Al
3+
and Fe
3+
ions on the fluoride mineralization
and their site preferences in the calcium silicate phases of Portland cement are investigated. It is
found that the reactivity of Portland cement and its strength properties may be improved
substantially by the addition of fluoride, Al
2
O
3
and Fe
2
O
3
in appropriate amounts.
A comprehensive study of the incorporation of fluoride ions in the calcium silicate
hydrate (CSH) phases of hydrated Portland cement has been conducted. This work provides
important structural information about the fluoride guest ions in the CSH, contributing to a
further understanding of the retarding effect of fluoride on the hydration of Portland cement.
Furthermore, a new solid-state NMR pulse sequence (Forth and Back Cross Polarization) has
been developed during this work. This experiment provides a selective detection of
19
F
resonances that are dipolar coupled to
29
Si which is used to investigate
19
F
29
Si connectivities in
the CSH structure.
Finally, AlOSi connectivities in the open framework structures of alkali-activated
materials and strtlingite (2CaOAl
2
O
3
SiO
2
8H
2
O) have been investigated using
29
Si{
27
Al}
REAPDOR experiments. These materials are of interest in recent cement research since they
have the potential of being alternative binders in concrete.


Resum
Den stigende anvendelse af alternative rmaterialer og brndstoffer i moderne Portland
cement produktion tilfrer en signifikant mngde af de skaldte fremmede komponenter, ssom
SO
3
, P
2
O
5
og fluorider, i den fremstillede cement. Disse oxider kan fremme dannelsen af
cement-klinker faserne og deres hydrauliske egenskaber, men de kan ogs have en negativ effekt
ved at hmme dannelsen af alit, som er den vigtigste hydrauliske komponent i Portland cement.
Dette Ph.d. projekt har to hovedforml, hvor det frste sigter mod en 30 % erstatning af
Portland cement-klinker med alternative materialer, som har hydrauliske egenskaber, uden at
reducere cementens styrke. Dette opns bl.a. ved at optimere klinkernes kemiske
sammenstning, sledes at der opns en forget reaktivitet for alit-fasen. Det andet ml i
projektet har vret at udvikle og implementere faststof dobbelt-resonans kerne magnetisk
resonans (NMR) eksperimenter til strukturkarakterisering af cement-materialer. I projektet er der
bl.a. blevet anvendt Cross Polarization (CP), Rotational-Echo DOuble-Resonance (REDOR) og
Rotational-Echo Adiabatic-Passage DOuble-Resonance (REAPDOR) NMR eksperimenter til
strukturkarakterisering af gst ioner i calciumsilikat-faserne (alit og belit) i Portland cement.
I studiet af optimeringen af Portland cement-klinkers kemiske sammenstning har der
vret srlig fokus p mineraliseringsegenskaberne af calciumfluorid. Mineraliseringseffekten af
fluorid-ionerne er blevet undersgt med forskellige faststof
19
F,
27
Al og
29
Si NMR teknikker.
Disse undersgelser har ogs inkluderet indflydelsen af trivalente metal-ioner (ssom Al
3+
og
Fe
3+
) p flourid-mineraliseringsprocessen og deres strukturelle omgivelser i calciumsilikat-
faserne. Hoved-konklusionen er, at tilstningen af fluorid, Al
2
O
3
og Fe
2
O
3
i passende mngder
kan signifikant forge Portland cements reaktivitet og trykstyrke. PhD afhandlingen inkluderer
ogs et omfattende studium af indbygningen af fluorid-ioner i calcium-silikat-hydrat (CSH)
fasen dannet ved hydratisering af Portland cement. Resultaterne herfra bidrager med vigtig
strukturel information om fluorid-ionernes placering i CSH fasen og afslrer mekanismen for
den retarderende effekt af fluorid-ioner p hydratiseringen af Portland cement. I disse studiet
blev der udviklet et nyt faststof NMR eksperiment kaldet Forth and Back Cross Polarization.
Dette eksperiment blev anvendt til at studere SiF sammenkdninger i CSH stukturen. Til
sidst er der foretaget en karakterisering af AlOSi netvrkstrukturerne i alkali-aktiverede
materialer og i den uordnede struktur for strtlingit (2CaOAl
2
O
3
SiO
2
8H
2
O) ved hjlp af
29
Si{
27
Al} REAPDOR eksperimentet. Alkali-aktiverede materialer har i jeblikket en stor
forskningsmssig bevgenhed, da de potentielt delvist kan erstatte Portland cement-klinker,
primrt med henblik p at reducere CO
2
emissionen fra cement produktionen.


Table of content



Futurecem ..1
Chapter 1: Cement Chemistry ................................................................................................. 3
1.1. Portland cement manufacture .................................................................................. 4
1.1.1. Quarrying and preparing of raw materials ............................................................... 4
1.1.2. Clinker production in the rotary kiln ....................................................................... 5
1.1.3. Cement grinding ...................................................................................................... 6
1.2. The main constituent phases of Portland clinker ..................................................... 8
1.2.1. Tricalcium silicate ................................................................................................... 8
1.2.2. Dicalcium silicate .................................................................................................... 9
1.2.3. Tricalcium aluminate, 3CaOAl
2
O
3
....................................................................... 10
1.2.4. Ferrite, Ca
2
(Al
x
Fe
1-x
)O
5
......................................................................................... 11
1.3. Minor components ................................................................................................. 12
1.3.1. Flux ........................................................................................................................ 13
1.3.2. Mineralizers ........................................................................................................... 14
1.4. Hydration chemistry for Portland cement ............................................................. 14
1.4.1. Hydration of calcium silicate phases ..................................................................... 14
1.4.2. Hydration of tricalcium aluminate ......................................................................... 15
1.4.3. Hydration of Portland cement ............................................................................... 16
1.5. Structure of the main hydration product CSH ................................................... 18
1.5.1. Tobermorite 14-, Ca
5
Si
6
O
16
(OH)
2
7H
2
O ............................................................ 19
1.5.2. Jennite, Ca
9
Si
6
O
18
(OH)
6
8H
2
O .............................................................................. 20
1.5.3. Portlandite, Ca(OH)
2
............................................................................................. 22

Chapter 2: Applications of solid-state NMR in cement research ....................................... 23
2.1. NMR theory ........................................................................................................... 24
2.1.1. Chemical shift interaction ...................................................................................... 25
2.1.2. Quadrupolar coupling interaction .......................................................................... 26
2.1.3. Dipolar coupling interactions ................................................................................ 27
2.2. Solid-state NMR techniques in cement research ................................................... 29
2.2.1.
29
Si MAS NMR ..................................................................................................... 29

2.2.2.
27
Al MAS NMR ..................................................................................................... 32
2.2.3.
19
F MAS NMR ...................................................................................................... 32
2.2.4.
43
Ca MAS NMR .................................................................................................... 34
2.2.5. Inversion-Recovery (IR) MAS NMR .................................................................... 36
2.2.6. Cross Polarization (CP) ......................................................................................... 37
2.2.7. Forth and Back Cross Polarization (FBCP) ........................................................... 38
2.2.8. Rotational-Echo Double-Resonance (REDOR) .................................................... 41
2.2.9. Rotational-Echo Adiabatic-Passage Double Resonance (REAPDOR) ................. 42
2.2.10. Multiple-Quantum (MQ) MAS NMR ................................................................... 46

Chapter 3: Fluoride Mineralization ...................................................................................... 49
3.1. Site preferences of F

in the calcium silicate phases of Portland cement ............. 50


3.1.1.
19
F MAS NMR ...................................................................................................... 50
3.1.2.
29
Si{
19
F} and
27
Al{
19
F} CP/MAS NMR ............................................................... 51
3.1.3.
29
Si{
19
F} CP-REDOR NMR.................................................................................. 53
3.2. Mineralizing effects of calcium fluoride and calcium sulphate ............................ 57
3.3. Secondary effects of fluoride mineralization......................................................... 60
3.4. Incorporation of Fe
3+
ions in the calcium silicate phases ...................................... 65
3.5. Hydration of fluoride-mineralized Portland cement .............................................. 71
3.6. Application of fluoride mineralization .................................................................. 76
3.6.1. Preparation of test clinkers .................................................................................... 76
3.6.2. Strength performances of fluoride-mineralized Portland cement.......................... 77
3.7. Summary ................................................................................................................ 79

Chapter 4: Fluoride-ion environments in CSH ................................................................ 81
4.1. A brief description of the CSH structure from the T/J viewpoint. .................... 82
4.2. Fluoride-ion environments in synthetic CSH and hydrated Portland cement ... 84
4.2.1. Site preferences of F

ions in the CSH structure from


19
F MAS NMR ............ 86
4.2.2. Influence of fluoride ions on the CSH structure ............................................... 89
4.3. Influence of fluoride ions on the hydration of Portland cement ............................ 95
4.3.1. Identification of CaF
2
in hydrated Portland cement by
19
F MAS NMR ............... 95
4.3.2. Hydration of fluoride-mineralized Portland cement from
19
F MAS NMR ........... 96
4.3.3. Retarding mechanism of fluoride ions on the hydration of Portland cement ...... 101
4.4. Summary .............................................................................................................. 103


Chapter 5: Framework structures of alumino silicates from NMR ................................. 105
5.1. SiOAl connectivities in alkali-activated materials .......................................... 106
5.1.1. Effects of Si/Al and Na/Al molar ratios .............................................................. 108
5.1.2.
29
Si {
27
Al} REAPDOR NMR .............................................................................. 111
5.2. Disorder in the double tetrahedral layer structure of Strtlingite ........................ 115
5.2.1.
29
Si MAS NMR ................................................................................................... 116
5.2.2.
29
Si{
27
Al} REAPDOR NMR ............................................................................... 117
5.3. Summary .............................................................................................................. 119
Conclusions .........120
Reference .123
Appendix 1 Sample preparations....143
Appendix 2 NMR Measurements and other analytical techniques 146
Paper I Site Preferences of Fluoride Guest Ions in the Calcium Silicate Phases............151
Paper II Incorporation of Fluoride Guest Ion in the Calcium Silicate Phases.....155
Paper III Characterization of Guest-Ion Incorporation...163
Paper IV Characterization of Guest-Ion Incorporation (in russian)...163
Paper V Site Preferences of F

, Al
3+
and Fe
3+
Guest Ions in the Calcium Silicate..173
Paper VI Characterization of the network structure of alkali-activated aluminosilicate..181
Manuscript I Characterization of the aluminosilicate network..191
Manuscript II The distordered structure of Strtlingite...........225
Manuscript III Fluoride mineralization of Portland cement..251


Introduction 1
FUTURECEM
Sustainable cement for the future

Cement is the essential glue in concrete, which is todays fundamental building material
with an annual consumption only surpassed by water. The cement manufacturing process is
however associated with a severe CO
2
emission. With an annual worldwide cement production
of about 3.3 billion tonnes in 2010
[1]
, the cement industry is responsible for about 5 % of the
man-made CO
2
emissions
[2-4]
. In fact, according to the increasing demand for concrete, as
growth and modernization take place in the developing countries, the cement production is
forecasted to double by the year 2050. Thus, identifying technology to reduce the CO
2
emission
intensity from cement production is urgent. The Getting the Number Right (GNR) Database
System
[5]
, which has registered data from 844 cement installations worldwide from 1999 to 2008,
reveals that about 60 % of the direct CO
2
emission originates from the chemical reactions during
cement clinker formation, e.g. the calcination process, CaCO
3
CaO + CO
2
. The fuel
combustion to reach temperatures of 1450 C in the cement kilns, necessarily for the formation
of clinker phases, is responsible for the remaining 40 %. Indirect emissions (e.g., from electric
power consumption) only contribute about 10 % to the overall CO
2
emissions. Based on current
knowledge, the International Energy Agency (IEA) and the World Business Council for
Sustainable Development (WBCSD) Cement Sustainability Initiative (CSI) have recently
presented the Cement Technology Roadmap 2009
[6]
, focusing on four distinct reduction levers
available in cement production: (i) thermal and electric efficiency, (ii) alternative fuel use, (iii)
clinker substitution and (iv) carbon capture and storage (CCS). The roadmap also emphasizes
that none of the options alone can yield the necessary reduction in CO
2
emissions. However,
from an economical viewpoint, option (i) is less attractive since an optimization of the energy
efficiency usually requires installation of new plants or upgrading of the old plants. CCS is a
relatively new technology and it has not yet been utilized at the industrial scale in cement
production. Thus, most of the ongoing developments and innovations focus on the use of
alternative fuels and clinker substitutions by other minerals with hydraulic and/or pozzolanic
properties
[7,8]
; those are usually referred to as Supplementary Cementitious Materials (SCMs).
However, the use of SCMs is complicated by several factors, e.g. high water demand, poor
workability, retention and a significant reduction in the early strength of the blended cement.
Therefore, the clinker substitution is typically limited to an average clinker factor of 0.78, i.e., 22
% of the clinker is replaced by SCMs including gypsum
[6]
.
Introduction 2
The FUTURECEM project was established in 2007 for a four year period (2007 - 2010)
with the main purpose to develop new building materials produced from readily available Danish
raw materials of low cost and low energy consumption, which should have the same
performance as todays concrete materials. The project is a joint collaboration between Aalborg
Portland A/S, iNANO at the Aarhus and Aalborg University, and the Geological Survey of
Denmark and Greenland (GEUS), with partly financial support from the Danish National
Advanced Technology Foundation. The first principal goal is to obtain a 30 % replacement of
the clinkers in concrete production by SCMs, resulting in an approx. 30 % reduction of the total
CO
2
emission. On a longer time scale, the FUTURECEM project also targets a larger
replacement of clinkers utilizing geopolymeric materials, such as alkali-activated metakaolin,
which possibly may result in a 70-80 % CO
2
reduction.
The research program for this PhD project is a part of FUTURECEM and focuses on
modifications of the Portland cement clinker composition with the aim of increasing the clinker
reactivity, which may enable a higher degree of clinker replacement. Particularly, the
mineralizing effects of calcium fluoride, which is widely used in todays Portland clinker
production, are investigated. The presence of fluoride ions has shown several important impacts
on the formation of the cement clinker phases as well as their hydrational reactivities
[9,10]
. The
optimization of the mineralizing properties for fluoride, which seems to be dependent on the
chemical composition of the clinkers, requires knowledge about the location of the fluoride ions
and their potential couplings to other guest ions in the anhydrous as well as hydrated phases of
Portland cement. Thus, another major goal of this project is to investigate the feasibility of solid-
state Nuclear Magnetic Resonance (NMR) spectroscopy in obtaining such information for
cementitious materials, which otherwise is difficult to be measured by other techniques.
The present dissertation consists of five chapters. The first two chapters give a brief
introduction to the relevant cement chemistry and the basic theory for NMR spectroscopy,
respectively. Chapter 3 is devoted to the investigation of fluoride mineralization of Portland
clinker while Chapter 4 concerns the influence of fluoride ions on the hydration of Portland
cement and their site preferences in the calcium silicate hydrate phases. Chapter 5 demonstrates
the potential applications of solid-state NMR in structural investigations of alumina-rich
cementitious systems such as alkali-activated materials and strtlingite. The dissertation includes
applications of NMR techniques such as Cross Polarization (CP), Rotational-Echo DOuble-
Resonance (REDOR) and Rotational-Echo Adiabatic-Passage DOuble-Resonance (REAPDOR)
on different nuclear-spin isotopes, including
19
F,
27
Al,
29
Si and
43
Ca, to obtain complementary
structural information for the studied cement phases.
Chapter 1. Cement Chemistry 3








1. Chapter



Cement Chemistry









Portland cement was patented by Joseph Aspdin in 1824
[11]
. The name refers to the
similarity in the color and strength properties of the cement and Portland stone, i.e. a limestone
quarried in Dorset, south west England. However, the mineralogy and hydrational properties of
Aspdins cement differ significantly from todays Portland cement. Nowadays, several cement
types are produced throughout the world for a wide range of applications, although the majority
has been developed for general constructional use
[12]
. Cements are named in accordance with
standard specifications, which define the chemical composition, physical properties and
performance requirements for the cement. The standard specifications also set the upper limits
for the content of each of the minor components. For example, the upper limit for the MgO
content in Portland cement is typically set to 5 %w because MgO levels above this limit result in
uncombined MgO (i.e. MgO that is not incorporated in the clinker phases), which can cause
destructive expansion in hardened concrete
[13]
. In fact, the standard specifications used by
countries around the world can be slightly different and therefore, various names may also be
given to the same class of cements
[14]
. For example, Portland cement in current European and
British standards corresponds to Types I and II Portland cement in the American standard, i.e
American Society for Testing and Materials (ASTM).
Chapter 1. Cement Chemistry 4
1.1. Portland cement manufacture
Portland cement is generally made from calcareous materials such as limestone,
marlstone or chalk. Furthermore, argillaceous materials including sand, clays or bauxite and iron
ores may be needed to provide additional SiO
2
, Al
2
O
3
and Fe
2
O
3
. Typically, the raw meal has a
chemical composition of 67 %w CaO, 22 %w SiO
2
, 5 %w Al
2
O
3
, 3 %w Fe
2
O
3
and 3 %w of
minor components such as MgO, SO
3
, alkali oxides and halogenides
[13]
. Various engineering
technologies are available for the Portland cement production. However, the short kiln with a
multistage of preheater and a cooling system is the current most efficient technology
[15,16]
. The
basic steps of the process will be briefly described below (for further reading see references [17
and 18]).

1.1.1. Quarrying and preparation of raw materials
Cement plants are often located close to naturally occurring calcareous and argillaceous
deposits to minimize the transport cost. The first step of the dry process involves crushing of the
quarried lime stone and clay into small pieces. In order to ensure a high reactivity for the raw
materials, they are milled together to produce a raw meal with more than 85 % of the particles
having a diameter below 90 m. The quality of the cement depends largely on the chemical
composition of the raw meal. Three important parameters
[13]
are widely used in clinker
production to design the chemical composition of the raw meal

) O Fe 0.65(%w ) O Al 1.2(%w ) SiO 2.8(%w
CaO %w
LSF
3 2 3 2 2
+ +
= (1.1)

3 2 3 2
2
O Fe %w O Al %w
SiO %w
SR
+
= (1.2)

3 2
3 2
O Fe %w
O Al %w
AR = (1.3)

The proportion of silicate phases in Portland clinker is determined by the silica modulus (SR)
while the lime saturation factor (LSF) mainly governs the alite (Ca
3
SiO
5
) to belite (Ca
2
SiO
4
)
ratio. For the pure CaOSiO
2
Al
2
O
3
Fe
2
O
3
system, a LSF of 1.0 or above indicates that the
formation of alite from belite and lime is saturated. For most productions, LSF is set in the range
Chapter 1. Cement Chemistry 5
of 0.92 0.98 to minimize the amount of uncombined CaO, i.e. the free CaO which are not
incorporated in the principal clinker phases. The alumina modulus (AR) is important for the
quantity of the clinker melt which is the only transport medium for the raw material in the rotary
kiln. In general, the silica and alumina modulus are in the ranges of 2.0 3.0 and 1.0 4.0,
respectively. However, this does not include special clinker types such as white Portland clinker
which has much lower aluminum and iron oxide contents. Variations in LSF, SR and AR can
significantly change the burnability, which is defined as the ease of combining belite and lime
into alite. The burnability may also be affected by several other factors, including the quantity of
the minor components and the grain size of the raw material. It is usually expressed by the
quantity of free CaO present in the final clinker.

1.1.2. Clinker production in the rotary kiln
The fine-grained raw meal is dried in a preheater tower, which consists of a series of
vertical cyclones, before it is fed into the rotary kiln. As the particles fall into the combustion
chamber, which is placed at the bottom of the preheater above the kiln, they have typically
reached a temperature of 800 900 C. In this temperature zone, approximately 90 % of the
limestone decomposes to lime in accordance with the chemical reaction
CaCO
3(s)
CaO
(s)
+ CO
2(g)
(1.4)
This process is usually referred to as calcination and is responsible for 60 65 % of the direct
CO
2
emissions from clinker production. The remaining ~40 % originates from the combustion of
fuel, which is fired directly into the discharge end of the kiln. As the tilted kiln slowly rotates,
about 3 5 rotations per minute, the calcined material slides and tumbles towards the flame. The
material is heated to partial fusion and several chemical reactions occur as a kiln temperature of
about 1450 C is progressively reached. Due to the rotation, the clinker leaves the kiln as
roughly spherical nodules of 1 3 cm in diameter. The relative proportions of clinker minerals
formed as the material moves into different kiln-temperature zones are graphically illustrated in
Figure 1.1. The chemical reactions for the clinker formation can be divided into three groups:
(1) The most important reactions occurring below ~1300 C are calcination, decomposition
of the clay minerals and formation of belite (2CaOSiO
2
), aluminate and ferrite. The
major phases at the end of this stage are belite, lime, aluminate and ferrite. Only a small
Chapter 1. Cement Chemistry 6
amount of clinker melt is formed at this temperature. The formation of belite, which is the
second most occurring phase in Portland clinker, can be expressed as
2CaO
(s)
+ SiO
2(s)
2CaOSiO
2(s)
(1.5)
(2) As the kiln temperature is raised above 1300 C, a melt phase is formed, consisting
mainly of tricalcium aluminate (3CaOAl
2
O
3
) and ferrite (Ca
2
(Al
x
Fe
1x
)
2
O
5
, where 0 < x
< 0.7). The pure CaO SiO
2
Al
2
O
3
Fe
2
O
3
system has a eutectic point, which occurs at
1338 C for an alumina modulus of AR = 1.38. A deviation of the alumina modulus from
this value requires much higher temperature to preserve the same level of clinker melt
content. In the temperature zone of 1300 1450 C, the clinker melt is the only transport
medium for belite and lime. Therefore, the quantity and properties (e.g., viscosity and
surface tensions) of the clinker melt are decisively important for the formation of alite.
Lime and belite are brought into reaction within the melt phase to form alite, in
accordance with the chemical reaction
CaO
(s)
+ 2CaOSiO
2(s)
3CaOSiO
2(s)
(1.6)
To some extent, the presence of the minor components such as SO
3
, P
2
O
5
, F and alkali
oxides may contribute to the quantity as well as the properties of the clinker melt. These
minor components originate from either the quarried raw materials or the combustion
fuel. However, they may also be deliberately added to the raw meal to modify the clinker
towards certain application requirements.
(3) The hot clinker falls into a cooler where it is immediately quenched either by incoming
air or by water. Upon cooling, the clinker melt re-crystallizes to aluminate and ferrite.
Depending on the cooling rate, different polymorphic transitions will also occur for alite
and belite.

1.1.3. Cement grinding
To produce cement, the clinker is mixed and finely ground with other hydraulic minerals
and usually 4 5 %w gypsum. Depending on the application requirements, the clinker content
can vary from 10 %w to 95 %w. For example, the clinker factor in blastfurnace cement CEM III
(C) is only 0.5 0.19, although this cement type is only used for a few special constructions
such as roads and tunnels.
Chapter 1. Cement Chemistry 7

Figure 1.1 Schematic illustration of the typical proportions of phases for the formation of Portland clinker minerals
as a function of the progressive kiln temperature. The figure is adapted from the work by Wolter
[19]
.

Chapter 1. Cement Chemistry 8
1.2. The main constituent phases of Portland clinker
The properties of Portland cement are mainly determined by the proportion of its four
principal clinker phases which are the impure forms of Ca
3
SiO
5
(alite), Ca
2
SiO
4
(belite),
Ca
3
Al
2
O
6
(tricalcium aluminate) and Ca
2
(Al
x
Fe
1-x
)
2
O
5
(ferrite). Other phases such as periclase
(MgO), quartz (SiO
2
), free lime (CaO), etc. may also be present in minor quantities, usually less
than 1 %w.

1.2.1. Tricalcium silicate
Alite is the most important constituent clinker component and typically constitutes 50
70 %w of Portland clinker. This highly hydraulic clinker mineral is an impure form of tricalcium
silicate, which can occur in seven crystal modifications: three triclinic
[20]
(T), three
monoclinic
[21]
(M) and one rhombohedral
[22]
(R) polymorphs. These alite structures are built of
Ca
2+
, O
2
and SiO
4
4
ions. They are all structurally similar with respect to the positions of Ca, Si
and O atoms, although they differ significantly in the orientation of the SiO
4
tetrahedra.
Furthermore, the mean coordination number of the Ca
2+
ions is different for the polymorphs. For
example, it is 5.66 in the R form, but 6.21 in T
I
polymorph. On heating, tricalcium silicate
undergoes a series of reversible phase transitions
[23]

R M M M T T T
C 1060
III
C 1050
II
C 990
I
C 980
III
C 920
II
C 600
I
o o o o o o

In fact, pure tricalcium silicate is only thermodynamically stable at temperatures above 1300 C.
On cooling, it becomes unstable between 1300 C and 1000 C with a maximum rate of
decomposition into belite and lime at 1175 C. At room temperature, pure tricalcium silicate is
meta-stable only in its T
I
form. The crystal structure for tricalcium silicate
[20]
in its triclinic form
is shown in Figure 1.2; the unit cell corresponds to the formula unit Ca
27
Si
9
(O
b
)
36
(O
i
)
9
,

where b
and i denote the oxygen sites involved in covalent SiO bonds and the interstitial oxygen sites,
respectively. The high-temperature polymorphs of tricalcium silicate can be stabilized at room
temperature by the incorporation of a sufficient amount of impurities in its structure
[24-27]
. For
example, alite is stabilized as the M
III
polymorph by the incorporation of Mg
2+
ions
[26]
in the
Ca
2+
sites while the rhombohedral form is stabilized by Zn
2+
incorporation
[28]
. The impurities are
often referred to as substituent oxides and their ions as guest ions. In Portland clinkers, alite
occurs mainly in the M
III
and/or M
I
forms, containing roughly 4 %w substituent oxides
[29-31]
. The
Chapter 1. Cement Chemistry 9
typical chemical compositions, including the most common encountered substituent oxides, for
Portland clinkers are summarized in Table 1.1.



Figure 1.2 Graphical illustration of the crystal structure for the triclinic form of tricalcium silicate. The unit cell
includes 9 non-equivalent SiO
4
-tetrahedra, 9 interstitial oxygens (O
i
) sites and 36 oxygens (O
b
) bonded directly to
the Si atoms. (-) Ca
2+
, (-): interstitial oxygen O
2
and (blue tetrahedra): SiO
4
4-
. The crystal structure data is adapted
from reference [20].


1.2.2. Dicalcium silicate
The second major constituent phase of Portland clinkers is the impure form of dicalcium
silicate, denoted belite. Its content in Portland clinkers is typically 5 30 %w. The structure of
dicalcium silicate is built of Ca
2+
and SiO
4
4+
ions and can occur in five different polymorph
modifications: one , one | and three o forms
[32]
. At room temperature, dicalcium silicate occurs
mostly in the form. On heating, the mineral undergoes phase transitions to form the high
temperature polymorphs

C 1425
'
High
C 1160
' '
Low
C 680 630 C 500
o o o o

<

Chapter 1. Cement Chemistry 10
The structure of -Ca
2
SiO
4
, which is similar to that of olivine, (Mg,Fe)
2
SiO
4
, is much less dense
as compared to the other polymorphs. Therefore, the transition from the high temperature
polymorphs to on cooling causes the sintered material to crack and fall to a more voluminous
powder. The transition | can be avoided by the incorporation of a sufficient amount of guest
ions in the dicalcium silicate structure. It can also be prevented if the |-Ca
2
SiO
4
crystallites are
sufficiently small. The arrangements of Ca
2+
and SiO
4
4+
ions in the high temperature phases are
very similar. However, they differ significantly from the form. Clinker belite typically contains
4 6 %w substituent oxides and occurs largely in the | form
[29-31]
. An illustration of the |
dicalcium silicate crystal structure is shown in Figure 1.3; the asymmetric unit contains a single
tetrahedral SiO
4
site and two distinct Ca
2+
sites, which have seven and eight oxygen atoms
within the first coordination sphere.


Figure 1.3 Graphical illustration of the crystal structure for the beta form of dicalcium silicate
[33]
. The asymmetric
unit only includes a single SiO
4
-tetrahedral site and two distinct Ca
2+
sites. (-): Ca
2+
and (blue tetrahedra): SiO
4
4-
.


1.2.3. Tricalcium aluminate
Pure tricalcium aluminate (3CaOAl
2
O
3
) has a cubic structure and does not exhibit
polymorphism
[34]
; it is built of Ca
2+
ions and rings of six AlO
4
tetrahedra. However, a rather
large amount of guest ions can be incorporated in the Ca
2+
as well as the Al
3+
sites, resulting in a
series of different structures
[35]
. In general, the tricalcium aluminate phase of Portland clinker
exhibits cubic and/or orthorhombic structures. These impure forms contains up to 13 %w and 20
%w substituent oxides, respectively. The most important guest-ion incorporations are Si
4+
and
Fe
3+
substituted in the tetrahedral Al
3+
sites. A large amount of other guest ions such as Na
+
, K
+

and Mg
2+
may also be incorporated in the Ca
2+
sites. The tricalcium aluminate content in
ordinary Portland clinker is in the range of 5 10 %w whereas it is much lower in white
Portland clinkers.
Chapter 1. Cement Chemistry 11
1.2.4. Ferrite
The ferrite (Ca
2
(Al
x
Fe
1-x
)O
5
) content in ordinary Portland clinker is usually 5 15 %w,
but much lower in white Portland clinker. The grey color of ordinary Portland cement is mainly
due to the presence of this iron-rich phase. In clinkers, ferrite is typically found closely mixed
with tricalcium aluminate. Due to the similarity in the cell parameters for the tricalcium
aluminate and ferrite phases, oriented intergrowth can occur for these phases on re-
crystallization from the clinker melt
[36]
. The structure of ferrite can be derived from that of
perovskite (CaTiO
3
). It can occur with any of the compositions in the solid-solution series
Ca
2
(Al
x
Fe
1-x
)
2
O
5
, where 0 < x < 0.7
[13]
. Each Ca
2+
coordinates to seven oxygen atoms.
Furthermore, Al
3+
and Fe
3+
are both distributed between octahedral and tetrahedral sites. Clinker
ferrite contains about 10 %w of substituent oxides, mainly SiO
2
and MgO. Other ions such as
Na
+
, K
+
and S
6+
are also present in minor concentrations. The guest ions are mainly incorporated
by substituting for either the Al
3+
or Fe
3+
sites whereas Ca
2+
exhibits almost no substitution
[29,37]
.


Table 1.1 Typical chemical compositions for the four principal clinker phases of Portland cement
[13]
.
CaO SiO
2
Al
2
O
3
Fe
2
O
3
MgO Na
2
O K
2
O SO
3
P
2
O
5
TiO
2
Mn
2
O
3
Alite 71.6 25.2 1.0 0.7 1.1 0.1 0.1 0.1 0.1 0.0 0.0
Belite 63.5 31.5 2.1 0.9 0.5 0.1 0.9 0.3 0.1 0.2 0.0
Aluminate 56.6 3.7 31.3 5.1 1.4 1.0 0.7 0.0 0.0 0.2 0.0
Ferrite 47.5 3.6 21.9 21.4 3.0 0.1 0.2 0.0 0.0 1.6 0.7

Chapter 1. Cement Chemistry 12
1.3. Minor components
In addition to CaO, SiO
2
, Al
2
O
3
and Fe
2
O
3
, Portland clinker contains approximately 3
%w of other components such as MgO, SO
3
and alkali oxides
[13]
; those are usually referred to as
minor component in cement chemistry. They originate mainly from the quarried raw materials
and the combustion fuel. The most encountered minor components in Portland clinker can be
divided into two groups according to their general influences on the clinkering process
[24,25]
:
(1) Flux agents include compounds that modify the temperature of liquid formation, the
properties of the clinker melt and the crystal morphology of the principal clinker
minerals.
(2) Mineralizer agents include compounds that influence the thermodynamic stability of the
principal clinker minerals. However, in many cases, it is only used to categorize the
compounds that influence the reaction of alite formation, i.e. Ca
2
SiO
4
+ CaO Ca
3
SiO
5
,
by modifying the thermodynamic properties of one or both calcium silicates.
Furthermore, some minor components can modify the hydration properties of Portland cement to
a significant extent. They typically involve one of the processes:
(1) Introduction of structural defects by forming solid solution within the principal clinker
mineral phases.
(2) Stabilization of clinker minerals in their high temperature polymorphs.
(3) Effects arising during cement hydration such as coating of the cement grains by insoluble
basic salts.
In general, the first two effects increase the hydrational reactivity of the clinker phases whereas
the last prevents the hydration of the anhydrous cement phases. Another important distinction
that also requires careful attention is the volatility of the minor components. Appreciably volatile
components typically have a higher concentration in the kiln atmosphere than that in the raw
material. Furthermore, they tend to be deposited in the cooler part at the end of the kiln due to
the recycling of air to reduce heat loss. For laboratory synthesis, the volatile components have a
rather high loss on ignition and thus, their contents in the clinkers can be much lower than that
added to the raw meal. SO
3
, K
2
O, Na
2
O, ZnO, and Cl
2
are very volatile while F, V
2
O
5
and As
2
O
3

Chapter 1. Cement Chemistry 13
have low volatility. Components such as MgO, P
2
O
5
, TiO
2
and NiO are essentially non-volatile
in the clinkering process.

1.3.1. Flux
Since almost all of the clinker melt is formed immediately at the eutectic temperature, the
reactants, e.g. lime and belite, are limited within local volumes by various proportions.
Therefore, the ease of combining lime and belite to form alite is strongly controlled by the
transport of the reactants through the clinker melt between such local volumes. The transport rate
is dependent on the quantity, viscosity and surface tension of the clinker melt. Most of the minor
components, when present in the raw meal, decrease the temperature for the first liquid
formation and contribute to the quantity of the clinker melt by approximately the same amount
as their content. Their influence on the viscosity and surface tension of the clinker melt is
somewhat complicated. In general, ions of strongly electropositive elements such as alkali metal
increase the viscosity while those of electronegative elements like Cl

and F

have a reverse
effect
[25]
. The influence of the p-block elements on the surface tension also follows a similar
trend, where the surface tension is decreased for an increase in the electronegativity. On the
other hand, the surface tension is increased as the electronegativity of the s-block elements is
increased.
The application of minor components for controlling the morphology of alite and belite
plays a very important role in clinker production. It is apparent from Table 1.1 that Mg
2+
, Al
3+

and Fe
3+
ions constitute the major part of guest ions in the calcium silicate phases. However,
compounds including ions such as S
6+
, P
5+
, K
+
and Na
+
have an increased interest in modern
Portland clinker production. For example, Mg
2+
ions are incorporated in the Ca
2+
sites and
stabilize alite in its M
III
polymorph
[26]
. S
6+
ions, on the other hand, substitute for the tetrahedral
Si
4+
sites and promote the formation of alite in its M
I
form. Furthermore, S
6+
ions, when
incorporated in belite, stabilize this phase in its | form
[38]
. Alkali ions and P
5+
are preferentially
incorporated in belite, stabilizing the o-belite polymorph, and they substitute for the Ca
2+
and
Si
4+
sites
[24,39,40]
, respectively. The guest ions may also compete for specific structural sites in
the clinker phases and therefore, limit or increase the substitution level for one another. For
example, Na
+
, K
+
and Mg
2+
ions substitute preferentially for the Ca
2+
sites while Al
3+
, P
5+
and
S
6+
ions compete for the tetrahedral Si
4+
sites.

Chapter 1. Cement Chemistry 14
1.3.2. Mineralizers
According to the restricted definition of a mineralizer which only includes compounds
that influence the reaction of alite formation
[25]
, the free lime content in the clinker can be used
to monitor the mineralizing effect. For example, the substitution of Si
4+
by Al
3+
ions in the
calcium silicate phases effectively increases the bulk SiO
2
content and therefore, increases the
belite content. However, this is a mass-balance effect and Al
3+
is not a mineralizer since the free
lime content remains low. On the other hand, S
6+
ion is a strong mineralizing agent on the
formation of belite. It increases the thermodynamic stability of belite by forming solid solutions
in this phase
[38]
. Therefore, the presence of a SO
3
source in the raw meal simultaneously
increases the amount of belite and the free lime content for the SO
3
-mineralized clinker. In
modern Portland cement production, CaF
2
and CaSO
4
are by far the most used mineralizers,
where F

strongly facilitates alite formation and SO


3
mineralizes belite
[41-45]
.


1.4. Hydration chemistry for Portland cement
The hydration of Portland cement is a process that includes many simultaneous chemical
reactions between the clinker minerals, gypsum and water, where the water/solid ratio required
in the cement mixture (paste) is typically 0.3 0.6 by weight
[13]
. Furthermore, the addition of
supplementary cementitious materials (SCM) such as natural pozzolans may also have important
impacts on the entire hydration process
[46-50]
. This section gives a brief description of the basic
hydration reactions for the clinker minerals as individual phases, which is fundamental for the
hydration of Portland cement in general. The description is based on references [13 and 51].

1.4.1. Hydration of calcium silicate phases
The main products formed from the hydration of alite and belite are calcium silicate
hydrates (CSH = xCaOSiO
2
yH
2
O), which constitute approx. 60 %w of the entire hydration
product of Portland cement. The relative proportions of phases formed from the hydration of
Portland cement are graphically illustrated in Figure 1.4. Typically, about 40 % of alite has
reacted within 1 day after mixing with water, 70 % within 28 days and virtually all in one year.
For a given grain-size distribution and water/solid (w/s) ratio, alite sets and hardens in a similar
manner to Portland cement. In this context, the term setting (or set) is used for stiffening without
Chapter 1. Cement Chemistry 15
increase in compressive strength while hardening (or harden) is the process of significant
strength development. Furthermore, hydration times up to 28 days are usually denoted early
hydration periods whereas the later hydration is referred to as long-term hydration. The
hydration of alite can be divided into three principle stages
[52]
:
(1) Within the first four hours, alite is attacked by water and a honeycomb-like product of
calcium silicate hydrates (CSH) is formed on the surface of the alite grains. Some
portlandite, i.e., Ca(OH)
2,
is also formed.

2 2 2 2 2
x)Ca(OH) (n O yH SiO xCaO O y)H x (n SiO nCaO + (1.7)
(2) During the middle stage, i.e. 4 24 hours, CSH and portlandite are rapidly formed.
The morphology of CSH formed at this stage strongly depends on the available space
for growth. Fibrous CSH has been observed as the outer product whereas the
honeycomb-like material grows in the surface fractures of the alite grains. As the
hydration proceeds after 24 hours, the inner product of CSH is also formed, which is
characterized by a very fine pore structure.
Belite follows the same hydration process as alite, however, with a much slower hydration rate;
approx. 30 % of belite has reacted during the early hydration period and about 90 % in one year.
Thus, the hydration of belite only contributes little to the early strength development, but it is
mainly responsible for the strength development during the long-term hydration.

1.4.2. Hydration of tricalcium aluminate
Tricalcium aluminate (C
3
A)
1
possesses a much higher reactivity as compared to alite and
belite. Initially, the meta-stable calcium aluminate hydrate phases, C
4
AH
13
and C
2
AH
8
, are
rapidly formed from the paste-liquid
[53]
. Subsequently, these phases are slowly converted to the
thermodynamically stable C
3
AH
6
phase. The overall hydration of C
3
A has finished within a few
hours after the mixing. This excessive hydration of C
3
A results in a severe stiffening, which
cannot be dispelled by remixing. This phenomenon is usually referred to as flash setting. In order
to retard the hydration of C
3
A, gypsum is normally added to the Portland clinker at the end of the

1
The nomeclature C = CaO, S = SiO
2
, A = Al
2
O
3
, F = Fe
2
O
3
and H = H
2
O is commonly used in cement chemistry.
However, it is only used to write the hydration reactions of tricalcium aluminate in this section and to abbreviate the
calcium silicate phases throughout the thesis.
Chapter 1. Cement Chemistry 16
clinkering process
[54]
. In the presence of sulphate ions, the C
3
A phase reacts rapidly to form
ettringite, which is an AFt phase.

32 3 6
Fast
2 3
H S A C 16H H S 3C A C + + (1.8)

12 4
Slow
32 3 6 3
H S 3C 4H H S A C A 2C + + (1.9)
When the sulphate ions are consumed by this reaction, ettringite slowly reacts with the
exceeding aluminate phase to form monosulphates (AFm), equation (1.9). The general formula
for the AFm (Al
2
O
3
F
2
O
3
mono) and AFt (Al
2
O
3
F
2
O
3
tri) phases are
Ca
2
(Al,Fe)(OH)
6
XwH
2
O and [Ca
3
(Al,Fe)(OH)
6
12H
2
O]
2
X
3
wH
2
O, respectively
[13,51]
. Here, X
in AFm denotes one formula unit of a singly charged anion such as OH

or half a formula unit of


a double-charged anion such as SO
4
2
or CO
3
2
. On the other hand, in AFt it denotes one formula
unit of a double-charged anion or two formula units of a single-charged anion. In addition to
setting regulation properties, the sulphate ions have important impacts on the hydration rate of
alite and the volume stability of the hydration products. The optimum content of gypsum,
however, depends on several factors, such as the cement composition, hydration time and other
conditions of hydration. The minimum content of SO
3
required to control setting is approx. 2
%w for an ordinary Portland cement.

1.4.3. Hydration of Portland cement
The microstructure of hydrated Portland cement develops in a similar manner to that formed
from the hydration of tricalcium silicate
[55,56]
. Soon after mixing, a gel-like layer containing
amorphous alumina, silica, calcium and sulphate in varying amounts is formed on the cement
grain surfaces. The AFt phase (e.g. ettringite) starts to precipitate within about 10 minutes after
mixing. After about four hours, the cement grains are completely covered by a thickening layer
of CSH. As the hydration proceeds, the CSH shells grow outward and inter-grow with
surrounding adjacent grains. This happens about 12 hours after mixing and usually it is referred
to as the cohesion point, where the setting has completed. A space filled with a highly
concentrated ion solution is observed between the shell and the inside of the anhydrous material;
the space can be up to 0.5 m wide. At this point, the CSH shell is still sufficiently porous and
ions can migrate between the inner and outer space through the shell wall. The hydration at this
Chapter 1. Cement Chemistry 17
stage is driven by dissolution and precipitation mechanisms. At a later stage (>24 hours), the
permeability of the shell is decreased significantly, preventing ion migration. CSH is formed
and deposits on the inside wall. The concentration of SO
4
2
ions inside the shells also drops
rapidly and AFt slowly reacts with the remaining aluminate phase to form AFm. The strength
development of Portland cement is mainly related to the formation of CSH, where the SiO
4

monomers released from the hydration of alite and belite polymerize into dimers and longer
chains of silicate tetrahedra.



Figure 1.4 Schematic representation of the hydration for the main Portland clinker phases and their resulting
hydration products. The areas of the boxes represent roughly the typical relative proportions of these phases as
reported in reference [13].

Chapter 1. Cement Chemistry 18
1.5. Structure of the main hydration product CSH
A calcium silicate hydrate (CSH) phase is the principal binding phases in hardened
cement
[13,51]
. The nanostructure of CSH exhibits several similarities to the structure of the
crystalline minerals tobermorite and jennite. However, CSH occurs with a wide range of
compositions, morphology and degree of structural order. The dashes indicate that no particular
composition is implied. The Ca/Si molar ratio in CSH formed from a nearly saturated alite
paste can possess a value in the range from about 1.2 to 2.1
[57]
. The average Ca/Si ratio is,
however, typically ~1.75
[57-59]
. An increase in the quantity of belite and/or SCMs results in a
decrease in the Ca/Si molar ratio, which typically enhances the long-term strength and durability
of the hydrated Portland cement, where the lower limit for the Ca/Si ratio in hydrated Portland
cement is 0.7
[57]
. Several models have been proposed for the nanostructure of CSH phases that
are formed from the hydration of calcium silicates, cement or related materials
[60-64]
. Many of
them exhibit several similar features, although they provide different degree of flexibility and
complexity. The models fall typically into two categories, which are derived from the structure
of tobermorite 14-. The first category, denoted T/CH, includes building blocks of tobermorite-
like structure (T) intermixed with layers of calcium hydroxide (CH). The second group, which is
usually denoted as T/J, contains elements of tobermorite and jennite (J) structures. The structures
of tobermorite, jennite and Ca(OH)
2
are described in Sections (1.5.1) (1.5.3), respectively.
Richardson and Groves introduced a generalized model for CSH
[62,65]
, which includes
formulations that could be interpreted from both T/CH and T/J structural viewpoints. The model
proposes the following general composition for the CSH

( ) ( )
{ } ( )
( )
O H Ca OH O Si H Ca
2
2
2 2 9 1 3 2
m
y n y n w n n w n

+
(1.10)
The number of silanol (SiOH) groups is given by w and the degree of protonation of the silicate
chains by w/n. The Ca/Si ratio can be obtained as

( )
( ) 1 3 2
4
Ca/Si

+
=
n
y n
(1.11)
The chemical formula for the main layer is given by the braces. The main layer contains 2n Ca
2+

ions surrounded by dreierketten silicate chains of mean length 3n 1. The charge balance is
preserved by n (w/2) interlayer Ca
2+
ions of the (ny)/2 given outside the braces. From the
T/CH viewpoint, the remainder of the (ny)/2 Ca
2+
ions occur in Ca(OH)
2
layers which
Chapter 1. Cement Chemistry 19
corresponds to the interlayer of tobermorite whereas, from the T/J viewpoint, they form a part of
the main layer resulting in regions with jennite-like structure. This flexible structure model
allows the Ca/Si ratio to vary from 1.0 to 2.5 by
(1) omission of part of the bridging SiO
4
tetrahedra in the dreierketten silicate chains,
resulting in an increased Ca/Si ratio. Ca
2+
may also be substituted into the bridging sites
leading to a further increase in the Ca/Si ratio,
(2) varying the degree of protonation, w/n; since, the increased Ca
2+
content can be balanced
by a decrease in the SiOH content and vice versa,
(3) incorporation of additional Ca
2+
ions can be charge-balanced by additional OH

. For the
T-based CSH structure the OH

ions are located in the interlayer whereas they are


incorporated in the main calcium layer of the J-based part.

1.5.1. Tobermorite 14-
Tobermorite 14-, Ca
5
Si
6
O
16
(OH)
2
7H
2
O, is the most hydrated phase of the tobermorite
mineral family
[66]
. The name refers to the basal spacing between the principal layers which
exhibit a chemical formula of [Ca
4
Si
6
O
16
(OH)
2
(H
2
O)
2
]
2
. The principal layer includes two
structurally nonequivalent calcium ions in sevenfold coordination and three tetrahedral Si
4+
sites.
The SiO
4
tetrahedra form dreierketten chains with periodicity of three silicon atoms, two Si on
the pair sites and one Si in the bridging site. The Ca atoms are coordinated to six oxygen atoms
and a hydroxyl group of the bridging SiO
4
tetrahedron forming a central Ca-O sheet, where the
dreierketten silicate chains are running on both sides. The CaO
7
and SiO
4
polyhedra form the
principal layer by sharing their vertices and apices as illustrated in Figure 1.5. The principal
layers are separated by an interlayer, which contains Ca
2+
, OH

and a large amount of water


molecules. The layers are held together by strong hydrogen bonds which are formed between the
oxygen atoms from the principal layer and the hydrogen atoms from water molecules present in
the interlayer.

Chapter 1. Cement Chemistry 20

Figure 1.5 Illustration of the layer structure for tobermorite 14-, Ca
5
Si
6
O
16
(OH)
2
7H
2
O. The calcium oxide in the
main layers is shown by green polyhedra. The dreierketten silicate chains shown in blue are formed from sequences
of three SiO
4
tetrahedra, one bridging and two pair sites. The interlayer consists of Ca
2+
and OH

ions, which are


depicted in grey and red, respectively. The crystal structure data is adapted from reference [66].

1.5.2. Jennite
The crystal structure of jennite
[67]
, Ca
9
Si
6
O
18
(OH)
6
8H
2
O, is classified in the space group
P-1. It exhibits a triclinic unit cell with a = 10.756 , b = 7.265 , c = 10.931 , o = 101.30,
| = 96.98 and = 109.65. The structure, Figure 1.6, is built of three distinct modules: ribbons
of two calcium octahedra sharing edges, dreierketten silicate chains and calcium octahedra on
inversion centers. The ribbons are connected to each other by sharing vertices. This results in a
zigzag layer of calcium layers, containing two types of ribbons: the first is sharing all its vertices
with other polyhedra while in the second ribbon, the apices on both sides of the zigzag layer are
Chapter 1. Cement Chemistry 21
corresponded to water molecules. The ribbons are further firmly linked through the dreierketten
silicate chains, which are present on both sides of the zigzag calcium layer. Furthermore, the
silicon atoms on the bridging sites are strongly coupled to the protons on water molecules. Both
calcium ribbons and silicate chains are running along the b axis. The composite octahedral-
tetrahedral layers are further connected through the additional calcium atoms in octahedral
coordination occurring on the inversion centers.


Figure 1.6 Illustration of the layer structure for jennite
[67]
, Ca
9
Si
6
O
18
(OH)
6
8H
2
O. The calcium oxides in the main
layers are shown by green polyhedra. The dreierketten silicate chains shown in blue are formed from sequences of
three SiO
4
tetrahedra, one bridging and two pair sites. The main layers are connected through the additional calcium
atoms in octahedral coordination occurring on the inversion center.
Chapter 1. Cement Chemistry 22
1.5.3. Portlandite
Ca(OH)
2
has a layer structure
[68]
(Figure 1.7) and belongs to the space group P-3m1. The
layers consist of one Ca
2+
sheet which is sandwiched by two sheets of hydroxyl groups (OH

).
Each calcium ion coordinates to six hydroxides forming a perfect octahedron with a CaO bond
length of 2.37 . Under ideal conditions, Ca(OH)
2
crystallizes with euhedral hexagonal shape. In
hydrated Portland cement, however, it exhibits a more massive and indeterminate shape.
Furthermore, a small amount of SiO
2
can be incorporated in the calcium layer. Ca(OH)
2

constitutes roughly 20 %w of the hydration products from Portland cement and this phase is
usually referred to as portlandite.


Figure 1.7 The layered crystal structure for Ca(OH)
2
. The calcium ions, shown in grey, are coordinated to six
hydroxyl groups. The oxygen atoms are shown in red and hydrogen atoms in blue. The crystal structure data is
adapted from reference [68].
Chapter 2. Applications of solid-state NMR in cement research 23









2. Chapter




Applications of Solid-State NMR
in Cement Research







The main advantage of solid-state NMR is related to its applicability in structural
investigations of crystalline as well as amorphous or poorly crystalline materials.
Complimentary to conventional analytical methods in cement research such as powder X-ray
diffraction (XRD), which generally detect the bulk or long-range order structures, different
NMR techniques can be used to obtain specific information about local structures. Generally, the
individual NMR-active spin isotopes exhibit different resonance frequencies and their NMR
interactions show a strong dependence on their specific site locations. Another important benefit
of NMR in cement research is its high sensitivity for specific NMR isotopes, making it possible
to investigate the structure of guest ions, which otherwise is very difficult by other techniques.
This Chapter gives an introduction to the basic solid-state NMR theory and its
applications in cement research. Furthermore, it includes relevant details for the NMR
techniques, which have been employed in this PhD project.
Chapter 2. Applications of solid-state NMR in cement research 24
2.1. NMR theory
For nuclear-spin isotopes with a non-zero spin-quantum number (I), when subjected to an
external static magnetic field (B
0
), their nuclear magnetic moments orient either parallel or
antiparallel with respect to the direction of ( )
0
B 0, 0, B = . The interaction between the nuclear
spin and the externally applied static magnetic field is the so-called Zeeman interaction, which
splits the degenerated energy states of the nuclear spin into (2I + 1) states. The detected NMR
resonance frequency (v) is related to the energy difference (E
2
E
1
) between these Zeeman
states by

eff
1 2
B
E E
=

= m

(2.1)
Here, is the Plancks constant (h/2t) and m is the magnetic quantum number, which can take
any of the values from I to I. In NMR, only single-quantum coherences, i.e. transitions with Am
= 1, can be detected directly. However, multiple-quantum coherences such as zero- and double-
quantum coherences of a two spin system or triple-quantum coherences for I 3/2 quadrupolar
nuclei may be indirectly detected in advanced NMR experiments. is the gyromagnetic ratio,
which is a specific constant for each of the nuclear-spin isotopes in the periodic table. B
eff
is the
effective magnetic field whose magnitude is determined by the externally applied magnetic field
and the local electronic and magnetic environments at the specific nuclear site. In essence, the
state of such a system is governed by quantum mechanics and can be fully described by the time-
dependent Schrdinger equation. However, solving the Schrdinger equation for a NMR
experiment, which often includes a very large number of spins (> 10
15
), can be rather difficult
and time consuming. Alternatively, the spin state in the NMR experiment is more conveniently
expressed by a spin-density matrix formalism using the Liouville-Von Neumann equation, which
in the absence of relaxation is given by
[69]

( ) ( ) ( ) | | t , t t
t
i H =
c
c
(2.2)

J D Q rf Z
H H H H H H H + + + + + =
(2.3)

where i is the complex imaginary unit. The density matrix, (t), is the statistical average for an
ensemble of spins. H(t) is the nuclear-spin Hamiltonian describing the external and internal spin
Chapter 2. Applications of solid-state NMR in cement research 25
interactions, where the external terms (H
Z
and H
rf
) are applied to modify the nuclear-spin state
whereas the internal interactions (H
Q
, H
o
, H
D
and H
J
) are nuclear-spin dependent and sensitive
to the local environment of the observed nuclear spin. H
Z
= v
L
I
Z
is the Hamiltonian for the
Zeeman interaction. For a given static magnetic field, B
0
, the Larmor frequency v
L
= B
0
is
specific for each nuclear-spin isotope. The H
rf
term in equation (2.3) represents the pulsed
radiation, applying an external rf field perpendicular to B, which in the laboratory-fixed frame
may be expressed as
( )
x carr 1 rf
I t cos 2 + = v v H (2.4)
where v
1
= B
1
/2t is the rf-field strength, v
carr
(~v
L
) is the carrier frequency of the pulse and
| is its phase relative to the direction of B
1
. The J-coupling (H
J
) is the direct spin-spin
interaction, i.e. a through bond coupling, which is usually small and has not been of importance
in the present study. A short description of the remaining internal interactions, i.e. the chemical
shift interaction (H
o
), dipolar couplings (H
D
) and the quadrupolar coupling interaction (H
Q
), is
provided below. In general, these interactions are very important in solid-state NMR of inorganic
diamagnetic solids.

2.1.1. Chemical shift interaction
The chemical shift interaction, including the isotropic chemical shift and chemical shift
anisotropy (CSA), originates from the chemical shielding of the external magnetic field caused
by the local electron distribution surrounding the observed nuclear spin. The Hamiltonian for the
chemical shift interaction between the nuclear spin I and the external magnetic field ( B
)
is
given by B I = o
o
H , where o is the chemical shift tensor, which in the Principal-Axis
System (PAS) can be expressed as
[70]


( )
( )
(
(
(

+

+
=
(
(
(

=
o
o
o
q o
q o
iso
iso
iso
zz
yy
xx
0 0
0 1 0
0 0 1
0 0
0 0
0 0

(2.5)
Here, the PAS elements fulfill the condition |o
zz
o
iso
| > |o
xx
o
iso
| > |o
yy
o
iso
| and the other
parameters are defined according to
Chapter 2. Applications of solid-state NMR in cement research 26
( )
zz yy xx iso

3
1
+ + = o
zz iso
= o

yy xx

=

(2.6)
o
iso
= o
iso
is the isotropic component, where the chemical shift (o) is positive to high frequency
(i.e. down-field) and the absolute shielding (o) is positive to low frequency (i.e. up-field). o
o
and
q
o
give the magnitude and the asymmetry of the chemical shift interaction, respectively.
Alternatively, the CSA can be characterized by the span (O = o
33
o
11
) and skew (k = 3(o
iso

o
22
)/O) parameters, where the PAS elements are chosen as o
11
> o
22
>

o
33
and o
iso
= 1/3(o
11
+
o
22
+ o
33
). In general, the NMR resonances are referenced by their chemical shift values quoted
in parts per million (ppm)

6
ref
ref sample
sample
10

= o (2.7)
This normalization of NMR resonances by standard reference samples (v
ref
) makes it possible to
compare data independent of the externally applied static magnetic field. In liquid-state NMR,
o
iso
is the only detectable component of the chemical shift interaction since the orientation-
dependent CSA is averaged out by molecular tumbling. A similar averaging can be achieved for
solids using Magic-Angle Spinning (MAS)
[71]
. In this technique, the powder sample is rotated
around an angle of 2 arctan = (~54.736) with respect to the external magnetic field. The
geometrical dependency of the CSA becomes time dependent, where the time-dependent terms
can be averaged out by employing a spinning frequency (v
R
) larger than the width of the
resonance in a static-powder experiment. For lower spinning speeds, the CSA results in a
characteristic spinning sideband (ssb) pattern with an envelope that resembles the lineshape in a
static NMR spectrum
[72]
; the positions of the spinning sidebands are separated by the value of v
R
.
The magnitude of o
o
and q
o
, which reflect the local electronic structure of the probed spins, can
be obtained from a simulation of the ssb pattern
[73]
.

2.1.2. Quadrupolar coupling interaction
NMR experiments for nuclear spin isotopes with I > 1/2 exhibit an additional interaction,
the quadrupolar coupling interaction ( I Q I
Q
=
+
H ), which is the coupling between the non-
spherical charge distribution of the nucleus and the electric-field gradients (EFGs) at the nuclear
Chapter 2. Applications of solid-state NMR in cement research 27
sites created by its surrounding electron density
[71]
. The quadrupolar coupling tensor in its PAS
is given as

( ) ( )
( )
( )
(
(
(

=
(
(
(

=
1 0 0
0 1 0
0 0 1
1 2 2
V 0 0
0 V 0
0 0 V
1 2 2
Q
Q
Q
zz
yy
xx
Q
h I I
C
h I I
eQ
q
q
(2.8)
where V(x,y,z) is the electric field gradient tensor which in its PAS fulfils |V
zz
| > |V
xx
| > |V
yy
| and
V
zz
= eq is used for the principal EFG tensor element. C
Q
= e
2
qQ/h is the quadrupolar coupling
constant, where eQ is the nuclear quadrupole moment, and q
Q
= (V
xx
V
yy
)/V
zz
is the EFG
asymmetry parameter (0 s q
Q
s 1.0). The quadrupolar coupling can possess a magnitude of
several MHz and thus, it cannot be averaged out by MAS. In general, it is sufficient to describe
the quadrupolar coupling interaction by the two lowest orders of its effective Hamiltonian
obtained by the Magnus expansion
[69]


~
) (
~
) (
~
H H H
2 1
Q Q
eff
Q + = (2.9)
The two terms are usually referred to as the first-order and second-order quadrupolar coupling
interactions. For quadrupolar nuclei with half-integer spin-quantum number, the first-order term
introduces a frequency modulation on the satellite transitions (m m 1, where m = 1/2)
resulting in a manifold of ssbs while it does not affect the central transition (1/2 1/2). The
first-order interaction can in principle be averaged out by MAS, although the achievable
spinning speed are often much to low for a complete averaging. On the other hand, all transitions
are affected by the second-order quadrupolar coupling interaction, which is only reduced by
MAS. For strong quadrupolar couplings, the centerband for the central transition exhibits a
characteristic lineshape with singularities, from which the quadrupolar coupling parameters C
Q
and q
Q
can be determined. The quadrupolar coupling parameters may also be obtained from a
simulation of the full spectrum of ssbs observed for the satellite transitions
[74]
.

2.1.3. Dipolar coupling interactions
The dipolar coupling is a through space interaction between two adjacent nuclear spins (I
and S), including two possible cases: (i) homo-nuclear dipolar couplings where I and S are
Chapter 2. Applications of solid-state NMR in cement research 28
identical spin nuclei and hetero-nuclear dipolar couplings where I and S are different spins. The
Hamiltonian for these interactions in a tensorial form is given as
[71]

S D I
D
=
+
H (2.10)
where the second-rank tensor for the dipolar coupling takes the following form in its PAS

(
(
(

=
(
(
(

=
d
d
d
0 0
0 2 0
0 0 2
D 0 0
0 D 0
0 0 D
D
zz
yy
xx
/
/
(2.11)
Here, d =
0
h
I

S
/(8t
2
r
3
) is the dipolar coupling constant, where
0
is the permeability in vacuum
and r is the internuclear distance for the IS spin pair. It is apparent from equation (2.11) that the
tensor is traceless with the largest PAS element (D
zz
) oriented along the dipolar coupling axis.
The absence of an isotropic component implies that dipolar couplings cannot be observed
directly in liquids.
In solid-state NMR, the first-order Hamiltonian for the dipolar coupling, when considered
as a perturbation to the Zeeman interaction, takes the form
( ) | | S I 3 S I 3 t
z z
D
=
D
H (2.12)
where z z y y x x S I S I S I S I + + = and v
D
is the frequency modulation caused by the dipolar
coupling interaction. For a pair of hetero-nuclear spins (i.e. I = S), the Hamiltonian can be further
reduced to
( )
z z
D
S I 2 t = H (2.13)
assuming that (v
L,I
v
L,S
) >> v
D
. This non-zero dipolar coupling interaction results in a
frequency modulation, which in the magic-angle spinning axis system is given by
( ) | | ( ) ( ) ( ) { } 2 sin t cos 2 sin t 2 cos
6 2

R
2
R D
+ +

=
d
(2.14)
Here, o and | are the azimuthal and polar angles describing the orientation of the internuclear
vector between I and S within the MAS frame. As the dipolar coupling constant is inversely
proportional to the cube of the distance between the spins (1/r
IS
3
), the condition of v
R
> v
D
is
Chapter 2. Applications of solid-state NMR in cement research 29
often achievable for hetero-nuclear dipolar couplings. However, the dipolar couplings between
high- spins such as
1
H and
19
F and an observed nuclear spin of lower Larmor frequency can be
very strong and therefore, they are not necessarily averaged out by MAS. Alternatively, different
homo- and hetero-nuclear decoupling sequences can be applied to eliminate or partly reduce the
line-broadening caused by the dipolar couplings on the observed resonances
[75-77]
.


2.2. Solid-state NMR techniques in cement research


2.2.1.
29
Si MAS NMR
29
Si is probably the most studied nuclear spin for cementitious materials, owing to the
high bulk SiO
2
content and the decisive role of the silicate phases in anhydrous as well as
hydrated cements. In general,
29
Si MAS NMR experiments are rather time-consuming due to the
low natural abundance of the
29
Si spin of 4.7 %. Furthermore, when present in cement phases,
the
29
Si spins may exhibit a long spin-lattice relaxation time
[78]
, which usually requires a
relaxation delay of 30 s at 7.1 T. From pioneering
29
Si MAS NMR studies of zeolites and
amorphous materials such as glasses and cements
[79-82]
, it has been demonstrated that the
29
Si
isotropic chemical shift for silicate mainly depends on the condensation of SiO
4
tetrahedra
(Figure 2.1), where an increased condensation (Q
i
, 0 s i s 4) corresponds to an up-field shift, i.e.
more negative o(
29
Si) value. In addition, the
29
Si isotropic chemical shift reflects the number of
Al atoms incorporated in the second-coordination sphere of the probed silicon sites, denoted
Q
i
(nAl, 0 s n si), for which each Si
4+
Al
3+
substitution leads to a down-field shift of
approximately 5 ppm. However, this implies that the chemical shift regions for different Q
i
units
overlap with each other, complicating the interpretation of the detected
29
Si resonances.
In accordance with the structure of triclinic tricalcium silicate determined by XRD, the
29
Si MAS NMR spectrum of a synthetic sample of this mineral identifies nine distinct silicon
environments with isotropic chemical shifts in the region for isolated SiO
4
tetrahedra (Q
0
)
[49]
,
Figure 2.2. For the monoclinic forms M
I
and M
III
, which are the commonly encountered forms of
alite in Portland cement,
29
Si MAS NMR shows broadened lineshapes from the overlapping
resonances of their 18 silicon sites
[83]
, covering the same spectral region. As it can be seen from
Figure 2.2, the M
I
and M
III
forms of alite are distinguishable by their characteristic
29
Si MAS
NMR lineshapes. In addition to the alite resonances, the
29
Si MAS NMR spectrum of anhydrous
Portland cement contains a narrow resonance at o(
29
Si) = 71.33 ppm from belite in its |
Chapter 2. Applications of solid-state NMR in cement research 30
form
[84]
. The isotropic chemical shifts for the Q
0
tetrahedra of the alite and belite structures can
be correlated with their corresponding mean SiO bond lengths, according to the linear
relationship
[85]

( ) ( ) 3 445 7 316
29
. . + =

d Si
O Si
o (2.15)




Figure 2.1 Schematic representation of the chemical shift regions for tetrahedrally coordinated silicon in
silicates
[79]
. Q
0
corresponds to the silicate monomers, Q
1
denotes the dimers and chain-end groups, Q
2
are the middle
groups in a silicate chain while Q
3
are chain-branching silicates and Q
4
are cross-linked framework silicates.

The fact that the
29
Si resonances for SiO
4
tetrahedra with different degrees of
condensation appear in distinguishable spectral regions can be utilized to follow the hydration
process for the silicate phases of Portland cement using
29
Si MAS NMR. Generally, the
hydration processes involve condensation of the Q
0
units to form dimer (Q
1
) and polymer (Q
2
)
silicates. The degrees of hydration for the individual phases (e.g. alite and belite) at each
hydration stage can be obtained by
Chapter 2. Applications of solid-state NMR in cement research 31
( )
( )
(

=
0
1
I
t I
t H (2.16)
where I
0
and I(t) are the normalized
29
Si intensities of the actual phase in the anhydrous and
hydrated samples, respectively.
The fraction of each individual phase (e.g. alite, belite and CSH) can be quantified
from a deconvolution of the
29
Si MAS NMR spectra
[83]
. In this project, the characteristic
lineshapes for alite in its monoclinic forms (Figure 2.2) have been simulated satisfactorily by
including nine
29
Si resonances. For the CSH phases covering the spectral region from 75
ppm to 90 ppm, the deconvolution includes two resonances at 76 ppm and 79 ppm for the Q
1
components and two resonances for the Q
2
(1Al) and Q
2
units at 81 ppm and 85 ppm,
respectively. Additionally, a resonance at 83 ppm has been included, accounting for protonated
Q
2
units, which have been identified for CSH samples cured in a CO
2
atmosphere (Chapter 4).


Figure 2.2
29
Si MAS NMR spectra (7.1 T and v
R
= 7.0 kHz) of synthetic samples of alite in the triclinic (a) and
monoclinic M
I
(b) and M
III
(c) forms. Nine resonances have been observed for the triclinic Ca
3
SiO
5
, for which the
rather broad peak at 69 ppm includes two overlapping resonances while the intensity of the peak at 73.44 ppm is
twice as large as the remaining resonances. The diamond (+) at 71.33 ppm indicates an impurity of | belite.
Chapter 2. Applications of solid-state NMR in cement research 32
2.2.2.
27
Al MAS NMR
27
Al MAS NMR is another very commonly applied tool in cement research. In contrast to
29
Si,
27
Al has a quadrupolar nuclear spin with the spin-quantum number I = 5/2, which in the
presence of an external magnetic field gives rise to six non-degenerated energy levels. The
27
Al
MAS NMR spectrum may be rather complex if all transitions are observed; the satellite
transitions of the
27
Al spin are substantially broadened by the first-order quadrupolar coupling
interaction. On the other hand, the central transition is only perturbed by the second-order
quadrupolar coupling interaction. Thus, in general, only the
27
Al MAS NMR spectrum for the
centerband of the central transition is considered in the analysis of the
27
Al MAS NMR spectra.
In ordinary Portland cement, the main part of the bulk Al
2
O
3
content is present in
tetrahedral coordination, which exhibits
27
Al resonances in the spectral region from about 40
ppm to 90 ppm at 14.1 T
[82,86]
. In general, the main aluminate phase (i.e. tricalcium aluminate)
appears as a broad resonance covering from about 40 ppm to 85 ppm. As a result of the high
natural abundance for the
27
Al spin of 100 %, it is also possible to detect Al
3+
guest ions that are
incorporated in the calcium silicate phases, despite their rather low concentration. The Al
3+
guest
ions from alite and belite appear as a rather narrow centerband at about 82 ppm with a shoulder
at 84 ppm, respectively. A small amount of Al
2
O
3
is also present in the ferrite phase, but these
27
Al spins cannot be observed due to their strong dipolar coupling with the electron spins of the
Fe
3+
ions.
Generally, the hydration of Portland cement involves a conversion of tetrahedral AlO
4

into octahedrally coordinated species (e.g. ettringite and monosulphate) and therefore, it may
easily be followed by
27
Al MAS NMR
[82]
. The
27
Al resonances from ettringite and monosulphate
appear at approx. 13.0 ppm and 9.0 ppm at 14.1 T, respectively. Furthermore, a resonance from
the tetrahedral Al
3+
guest ions in the CSH phases has been observed at about 60 ppm.

2.2.3.
19
F MAS NMR
The increasing application of fluoride mineralization in Portland cement production has
been the major motivation for the application of
19
F MAS NMR in structural investigations of
cementitious materials in this project. Although several
19
F MAS NMR studies are available in
the literature for fluoride structures in cement related materials such as glasses
[87-90]
, only a
single study on Portland cement
[91]
has been reported earlier.
19
F is a spin I = 1/2 nucleus with a
natural abundance of 100 %. Due to the high gyromagnetic ratio of the
19
F spin, close to the
Chapter 2. Applications of solid-state NMR in cement research 33
value of
1
H, the method is very sensitive, enabling fluorine to be detected even when present in
very small concentrations.
Previous studies of several different metal fluoride compounds and fluorine minerals have
demonstrated that the o(
19
F) chemical shift is strongly affected by the type of counter ions
[92]
,
and distributed over a spectral region of about 200 ppm with o-PbF
2
at high frequency (20
ppm) and NaF on the other low-frequency side at approximately 222 ppm. The o(
19
F) chemical
shifts for different SiF and AlF covalent bonds of silicon and aluminum in octahedral,
tetrahedral and five-fold coordination are predicted to appear in distinct spectral regions
[90]
.
However, the presence of different types of counter ions, such as Ca
2+
and Na
+
,

may introduce
additional frequency shifts to the observed
19
F nucleus, merging the chemical shift regions for
the different fluorine environments of the central cation coordination states. This is illustrated in
Figure 2.3 by
19
F MAS NMR spectra for fluoride with different counter ions.


Figure 2.3
19
F MAS NMR spectra (7.1 T, v
R
= 8.0 12.0 kHz) of NaF (a), SrF
2
(b), Na
2
SiF
6
(c) and MgSiF
6
(d).
Spectra (a) and (b) demonstrate the effect of counter ions on the chemical shift of fluoride ions while (c) and (d)
show the effect of counter ions on fluorine in covalent SiF bonds with the silicon atom in octahedral coordination
state. The asterisk shown in (a) indicates an artifact.
Chapter 2. Applications of solid-state NMR in cement research 34
2.2.4.
43
Ca MAS NMR
The application of
43
Ca NMR spectroscopy has been very limited since just about twenty
43
Ca NMR studies of organic as well as inorganic compounds have been reported
[93-101]
. Only
few of these studies concerns the application of
43
Ca MAS NMR in structural investigations of
Portland cement or its related materials
[93,94]
. The major difficulties in obtaining
43
Ca MAS NMR
spectra with an acceptable signal to noise (S/N) ratio are related to the low natural abundance of
the
43
Ca spin isotope (0.135 %) and its low gyromagnetic ratio, e.g. its Larmor frequency is only
40.39 MHz at 14.1 T
[95]
. Furthermore, since
43
Ca is a quadrupolar spin (I = 7/2), when present in
asymmetric environments such as in Ca(OH)
2
, its centerband from the central transition is
severely broadened by the strong second-order quadrupolar interaction. These difficulties are
illustrated in Figure 2.4 for a sample of Ca(OH)
2
. The
43
Ca MAS NMR spectrum was recorded
at 14.1 T using a 7.5 mm PSZ rotor (450 l sample volume). To get a decent signal to noise ratio
(S/N = 11.9), the experiment employed 196,608 scans with a relaxation delay of 2 s,
corresponding to a spectrometer time of 111 hours. The quadrupole coupling parameters (o(
43
Ca)
= 63 ppm, C
Q
= 2.65 MHz and q
Q
= 0.22) for the
43
Ca spins in Ca(OH)
2
were determined from a
spectral simulation of the second-order quadrupolar lineshape (Figure 2.4) using the STARS
software package
[74]
. This strong quadrupolar coupling leads to a broad resonance for the
centerband of the central transition, covering a region of nearly 40 ppm.


Figure 2.4
43
Ca MAS NMR spectra (14.1 T) of Ca(OH)
2
(a, v
R
= 5.0 kHz) and a synthetic sample of cuspidine
Ca
4
Si
2
O
7
F
2
(b, v
R
= 3.0 kHz). The spectra were obtained using a solid t/2-pulse of 2 s, a relaxation delay of 2 s
and 196,608 and 128,000 scans, respectively. The simulation of the second-order lineshape for the
43
Ca spins in
Ca(OH)
2
is shown below the spectrum (a).
Chapter 2. Applications of solid-state NMR in cement research 35
Another example, shown in Figure 2.4, is for a synthetic sample of cuspidine
(Ca
4
Si
2
O
7
F
2
), in which calcium is distributed among four distinct crystallographic sites. The
43
Ca MAS NMR spectrum (14.1 T, v
R
= 3.0 kHz) was acquired over 74 hours, i.e. 128,000 scans
for a relaxation delay of 2 s. The cuspidine sample was packed in a thin wall 7 mm Si
3
N
4
rotor
(310 l sample volume). This experiment results in a
43
Ca MAS NMR spectrum with S/N =
10.6. Due to the severe overlap of resonances, covering a spectral region from approximately
30 ppm to 30 ppm, in conjunction with the low S/N ratio, the different Ca environments in the
cuspidine structure cannot be distinguished.
Alternatively, for Ca sites with weak quadrupolar couplings such as in CaF
2
, the Cross-
Polarization sequence may be applied to enhance the sensitivity in the
43
Ca NMR experiment.
This is demonstrated in Figure 2.5 for a synthetic sample of CaF
2
. The standard single-pulse
43
Ca
MAS NMR spectrum was recorded with 17,280 scans, corresponding to 73 hours and the
spectrum has a S/N ratio of 58.8. However, a significant increase in the S/N ratio is obtained for
the
43
Ca{
19
F} CP/MAS spectrum, which has S/N = 140.3 and was recorded with only 6,144
scans (~26 hours).


Figure 2.5
43
Ca MAS (a) and
43
Ca{
19
F} CP/MAS NMR spectra (14.1 T, v
R
= 3.0 kHz) of a synthetic sample of
CaF
2
. The sample was doped with a small amount of NiO in order to reduce the spin-relaxation time for the
19
F and
43
Ca spins. The spectra were recorded using a 15-s relaxation delay and 17,280 and 6,144 scans, respectively. The
signal to noise (S/N) ratios in the spectra are 58.8 and 140.3, respectively. The vertical scale is expanded by a factor
of 11.6 in (a) relative to (b).
Chapter 2. Applications of solid-state NMR in cement research 36
2.2.5. Inversion-Recovery (IR) MAS NMR
A reliable quantification of the intensities in NMR spectra generally requires that the
spectra are recorded with a relaxation delay of at least d1 > 5T
1
, where T
1
is the spin-lattice
relaxation time. Determination of the magnitude for T
1
can be obtained by the Inversion-
Recovery (IR) pulse scheme
[102]
(Figure 2.6), where a systematic increase of the recovery time
(t) results in a series of spectra, for which the observed intensity is governed by the exponential
relation
[103]

( ) ( )
(

|
|
.
|

\
|
+ =
1
0
exp 1 1
T
t
M t M o (2.17)
Here, M
0
is the equilibrium magnetization, i.e. M(t = ), and o is a constant related to the pulse
imperfections, where o = 1 for ideal pulses.


Figure 2.6 Illustration of the Inversion-Recovery (IR) pulse scheme. Systematic increase of the recovery time (t)
results in a series of spectra, from which the T
1
or T
1

relaxation time can be determined from a fit of the


experimental data to equation (2.17) or (2.18).

The IR experiment can also be used to study paramagnetic ions indirectly. For dilute spin
systems, the spin-lattice relaxation time constant of the observed spin (I) may be dominated by
the strong dipolar interactions between the I spin and the free electrons spin (S) of the
paramagnetic ions. The inversion-recovery magnetization for such systems can be fitted to a
stretch exponential equation
[103,104]

( ) ( )
(
(

|
|
.
|

\
|
+ =
'
1
0
exp 1 1
T
t
M t M o (2.18)
where the spin-lattice relaxation time, usually denoted by T
1

,
which in the absence of spin
diffusion is related to the average concentration of paramagnetic ions (N
P
) by
Chapter 2. Applications of solid-state NMR in cement research 37

e I
e I I
2
P
3
1
1
1
5
2
, N
9
16 1
t v
t
t
+
+ = =
e
I
S S C C
T
) (
'
(2.19)

I
and v
I
are the gyromagnetic ratio and the Larmor frequency of the I spins while
e
and t
e
are
the gyromagnetic ratio and the spin-lattice relaxation time of the electron spins, respectively.

2.2.6. Cross Polarization (CP)
The CP experiment (Figure 2.7) involves a magnetization transfer from one spin (I) to
another (S) via the IS dipolar coupling obtained by matching the rf-field strengths applied on
the I and S spins
[105]
. In structural investigations of cementitious materials, the CP experiment
may be useful as a result of two main features
[106]
. First of all, according to the fact that the
dipolar coupling is inversely proportional to the cube of the internuclear distance (1/r
3
IS
), the CP
pulse sequence can be used as a filter which only detects IS spin pairs with internuclear
distances of usually less than 5 . Secondly, CP may be used to enhance the sensitivity for NMR
experiments of nuclei (S) with low gyromagnetic ratio, low natural abundance and/or long spin-
lattice T
1
relaxation times, such as
29
Si and
13
C
[107,108]
. In this case, the magnetization from an I
spin, typically with a high natural abundance, large gyromagnetic ratio and short T
1
relaxation
time (e.g.
1
H and
19
F), is transferred to the dilute S spin. The sensitivity for the detection of the S
spins in such an experiment is enhanced by a theoretical factor of
I
/
S
since the magnetization is
initiated from the I spins. Furthermore, as the relaxation delay depends on the T
1
relaxation time
of the I spins, a substantial reduction in the spectrometer time may be achieved.
The magnetization transfer requires that the rf-field strengths (v
1
) applied on the I and S
spins in the spin-lock period fulfill the Hartmann-Hahn (HH) matching condition
[109,110]
:

R 1S 1I
1 1 1 1 n ) m(m ) S(S ) m(m ) I(I + = + (2.20)
where n = 1, 2, v
R
is the spinning frequency, I and S are the spin-quantum numbers of the I
and S spins, and m are the magnetic-quantum numbers defining the transitions between the
Zeeman states (e.g. m m 1).
The HH-matching condition is simply reduced to v
1I
= v
1S
nv
R
for spin-systems where
both I and S are spin-1/2 nuclei. If quadrupolar nuclei (S > ) are involved in the spin system,
the CP transfer during the spin-lock period become rather complex since the satellite transitions
of S are substantially perturbed by the first-order quadrupolar interaction
[111]
. Moreover, MAS
Chapter 2. Applications of solid-state NMR in cement research 38
causes the quadrupolar frequency to oscillate back and forth between v
Q
. Depending on the
magnitude of v
Q
= 3C
Q
/[2S(2S 1)], the quadrupolar frequency experiences two or four zero-
crossings per rotor cycle. These oscillations result in a population interchange between the
2S + 1 Zeeman states of the quadrupolar nucleus and thereby, a modulation of the HH-matching.
Consequently, the CP transfer from the I to S spins is only efficient for the central transition of
the quadrupolar nucleus. Three regimes have been observed for the CP behavior during the spin-
lock period
[112]
: adiabatic passage (o >> 1), intermediate passage (o ~ 1) and sudden passage
(o << 1), with the o parameter defined by ) /(
R Q
2
1S
= . Efficient spin-lock can be achieved
both in the adiabatic and sudden passage regimes, while it is not possible for the intermediate
passage. For I = 1/2, the Hartmann-Hahn matching condition in the adiabatic and sudden passage
regimes are (i)
R 1 1I
) ( n S
S
+ = for v
Q
>> v
1S
and (ii)
R 1 1I
n
S
= for v
Q
<< v
1S
,
respectively.


Figure 2.7 Cross Polarisation (CP) pulse sequence. The dashed line illustrates the magnetization transfer from the I-
spins to the S-spins during the spin-lock period when the Hartmann-Hahn matching condition is fulfilled.

2.2.7. Forth and Back Cross Polarization (FBCP)
In practice, the standard CP experiment (Figure 2.7) requires that the initiated
magnetization originates from a highly abundant I spin with large , since the detection of the
reverse processes from a low- nucleus may require unreasonable long spectrometer time. For
spin systems where both I and S are low- spins, the magnetization can be initiated from a third
nuclear spin, typically
1
H, as employed in the Double Cross-Polarization (DCP) experiment
[113]
.
The DCP pulse scheme, which is a modification of the standard CP, consists of two cross
polarization periods: the first period is from the
1
H spins to one of the two low- nuclear spins
and the second is cross polarization between the low- nuclear spins.
Chapter 2. Applications of solid-state NMR in cement research 39

Figure 2.8 Forth and Back Cross Polarization (FBCP) pulse scheme. In this experiment, the initial
19
F
magnetization is transferred to
29
Si during the first CP period. This magnetization is stored along the x-axis and
transferred back to the
19
F spins by the second CP period. The removal of artifacts is obtained by phase-cycling the
pulses in accordance to the CYCLOPS method. |
1
= x x x x, |
2
= x x x x, |
3
= x x x x, |
4
= x x x x, |
rec
= x
y x y.

Figure 2.8 shows an alternative pulse scheme, which can be applied to obtain cross
polarization from a low- nuclear spin to a high- spin. The design of this pulse scheme is
inspired by the work of Wilhelm et al.
[114]
, who utilizes the
1
H
13
C dipolar coupling to study the
1
H diffusions for alanine and polyethylene in a
13
C{
1
H} 2D CP/MAS NMR experiment. The
Forth and Back Cross Polarization (FBCP) experiment was developed to selectively detect the
fluoride guest ions incorporated in the CSH phase (cf. Chapter 4). In analogy to the DCP
experiment, the FBCP experiment consists of two cross polarization periods, although only two
different types of nuclear spins are involved, e.g.
19
F and
29
Si. As in standard CP, the initial
magnetization is created by employing a t/2-pulse on the high- nuclear spin (e.g.
19
F).
29
Si
magnetization is built up during the first CP period and is being spin-locked along the x-axis for
the rest of the pulse scheme. The transverse
19
F magnetization remaining after the first CP period
is converted into longitudinal magnetization using a t/2 pulse which is phase shifted by 90
relative to the spin-lock rf field applied on the
19
F channel. A short delay (t) of typically 0.5 ms
is applied after the t/2 pulse to ensure a complete dephasing of the transverse
19
F magnetization.
During the second CP period, the magnetization is transferred back to
19
F from the
29
Si spins.
The HH-matching condition for the two CP periods is identical since the same
19
F
29
Si spin
pairs are involved. The experiment is further optimized using the phase cycle given in Figure
2.7.
The
19
F{
29
Si} FBCP/MAS NMR experiment has been tested on a powder mixture of a
synthetic sample of cuspidine and Teflon. Cuspidine is a rare natural abundant mineral with the
general formula of Ca
4
Si
2
O
7
F
2
[73,115]
, which has also been found in steelmaking slags. Its unit
cell consists of two crystallographically distinct SiO
4
sites, forming a Si
2
O
7

unit, and two


Chapter 2. Applications of solid-state NMR in cement research 40
distinct F

sites. Both F

ions are located in the vicinity of the Si


2
O
7

unit with SiF distances of


about 4.0 . On the other hand, Teflon is a polymer of tetrafluoroethylene with a general
formula of [CF
2
CF
2
]
n
. A single-pulse
19
F MAS NMR spectrum of this powder mixture is
shown in Figure 2.9 (a). The
19
F resonances originating from cuspidine appear at o(
19
F) = 101.6
ppm and o(
19
F) = 106.1 ppm while the
19
FC sites from Teflon exhibit a chemical shift of
o(
19
F) = 121.7 ppm. On the other hand, the
19
F{
29
Si} FBCP/MAS NMR spectrum recorded for
the same powder mixture only shows the
19
F resonances from the dipolar-coupled
19
F
29
Si spin
pairs of cuspidine, Figure 2.9 (b). In order to verify that these resonances are built up due to the
second
19
F
29
Si CP period, a control experiment has been conducted in which the rf-irradiation
on the
19
F spins during the second CP period is removed. In this case, the
19
F resonances from
Teflon as well as cuspidine are absent in the resulting spectrum, which strongly validates the
forth and back cross-polarization property of the
19
F{
29
Si} FBCP/MAS NMR experiment.



Figure 2.9
19
F MAS NMR (a) and
19
F{
29
Si} FBCP/MAS NMR (b) spectra of a powder mixture of a synthetic
sample of cuspidine and Teflon. The experiments were performed at 7.1 T using a spinning speed v
R
= 10.0 kHz, a
relaxation delay of 8.0 s, and 4 scans and 18571 scans, respectively. The single-pulse experiment allows
19
F
resonances from both cuspidine (101.6 and 106.1 ppm) and Teflon (121.7 ppm) to be detected. On the other
hand, the
19
F{
29
Si} FBCP/MAS NMR experiment provides a selective detection of the
19
F spins from the dipolar-
coupled
19
F
29
Si spin pairs of cuspidine.
Chapter 2. Applications of solid-state NMR in cement research 41
2.2.8. Rotational-Echo Double-Resonance NMR (REDOR)
Magic-angle spinning is a very important technique in solid-state NMR spectroscopy.
The spinning of the powder sample at the magic angle (54.736 ) can averages out several
anisotropic interactions including the CSA and hetero-nuclear dipolar interactions, providing
high-resolution NMR spectra. However, the anisotropic interactions may contain valuable
information about the local environment of the probed nuclear spins. Thus, several double-
resonance NMR pulse sequences
[116-118]
, including REDOR, have been developed with the aim
of retaining the anisotropic interactions and at the same time preserve the high resolution
provided by MAS.
If the condition of v
R
> v
D
is fulfilled, the modulated dipolar frequency given by equation
(2.14) becomes zero when integrated over a full rotor period (T
r
) of MAS. In the REDOR
experiment
[117,119]
, the t-pulses applied in the middle of each rotor cycle result in a sign-reversal
of the dipolar frequency in the following half-rotor cycle (Figure 2.10). Consequently, the
effective dipolar coupling becomes finite and gives rise to a dipolar dephasing of the
magnetization; the attenuated signal is denoted by S. The dipolar coupling between the nuclear
spins is monitored through a series of different experiments where the evolution period is
increased systematically in steps of nT
r
. To account for T
2
relaxation, a spin-echo experiment,
i.e. when the I-channel (
19
F in Figure 2.10) is turned off, is performed on the S-spins (
29
Si) to
produce the full signal (S
0
) for each evolution period. The REDOR fraction is obtained by
subtracting the attenuated signal from the full signal (AS = S
0
S). A plot of AS/S
0
as a function
of the evolution period (nT
r
) gives the universal REDOR curves with distinct slopes and
modulations, reflecting the dipolar coupling constant for the probed I-S spin pairs. The analytical
solution for the dipolar dephasing of an isolated spin-pair, in which both I and S are spin-
nuclei, can be expressed by an expansion using Bessel functions J
k
(x)
( ) ( ) ( ) ( ) d
S
S
sin cos sin nT 2 4 cos 1
r
0
= (2.21)
( ) | | ( ) | |

=
+ =
1 k
2
r k 2
2
r 0
0
nT 2 J
1 16k
1
2 nT 2 J 1 d d
S
S
(2.22)
Here, k is the order of the Bessel function
[120]
. The appearance of the REDOR curve becomes
very complicated when multiple-spin systems (I
n
S) are considered
[121]
. Furthermore, the
REDOR sequence is only appropriate for studying nuclei with spin-quantum number of 1/2 since
Chapter 2. Applications of solid-state NMR in cement research 42
the t-pulses are inefficient on the satellite transitions for quadrupolar spins
[122,123]
. The
quadrupolar coupling also introduces an oscillation of the REDOR fraction as a function of the
evolution time.
The REDOR experiment has been applied with success in structural investigations of a
wide range of crystalline as well as amorphous materials
[98,124-126]
. This experiment is frequently
used to identify and distinguish between independent I
n
S connectivities. Furthermore, it
emerges as a useful technique for determination of the IS distance of isolated two-spin systems.


Figure 2.10
29
Si{
19
F} CP-REDOR pulse scheme. The
29
Si
19
F dipolar coupling is re-introduced by a string of t-
pulses which are phase-cycled in accordance with the xy-4 sequence
[127,128]
, i.e., (xyxy)
n
. The t-pulses are separated
exactly by one half rotor period. The initial CP period
[108]
is applied to enhance the sensitivity of the experiment and
to selectively detect the
29
Si
19
F spin pairs.

2.2.9. Rotational-Echo Adiabatic-Passage Double Resonance (REAPDOR)
The REAPDOR pulse scheme
[129]
(Figure 2.11) is somewhat similar to the REDOR
sequence. However, this experiment has been developed for spin systems that involve
quadrupolar spins. In the experiment, it is utilized that the population of the 2S + 1 Zeeman
states can be interchanged due to the oscillations of the quadrupolar frequency during MAS, cf.
Section (2.2.6). This transfer process is facilitated by a single adiabatic-pulse of length t, which
should be shorter than one rotor period (T
r
) to keep the oscillations of the quadrupolar frequency
at a low number of zero-crossings
[122,130]
. In our studies, the REAPDOR experiments have
employed an adiabatic pulse of length T
R
/3, which maximizes the number of odd zero-crossings
for the quadrupolar frequency. Consequently, the oscillations caused by the quadrupolar
coupling are almost eliminated from the REAPDOR fraction. A recent study of the REAPDOR
sequence shows that the appearance of its AS/S
0
curve is determined by the spin-quantum
number of the probed quadrupolar nucleus
[122]
. In general, the dipolar dephasing, reflected by the
Chapter 2. Applications of solid-state NMR in cement research 43
REDOR and REAPDOR curves, becomes much complicated for multiple-spin systems.
However, for small dipolar dephasing, AS/S
0
s 0.3, the slope of the curves can be approximated
by a second-order polynomial
[131,132]

( )
( )
2
2
2
r
0

1
nT M
I I I S
S
+
= (2.23)
M
2
is the dipolar second moment, which reflects the strength of the dipolar coupling
( ) 1
1
4 15
4
I
6
IS
2
0 2 2 2
2
+
|
|
.
|

\
|
|
.
|

\
|
= I I
r

M
n
S I
(2.24)
where I is the spin-quantum number of the non-observed nuclear spin and n
I
is the number of I
spins that contribute significantly to the I-S dipolar coupling. The remaining parameters used in
equation (2.24) have their usual meanings according to standard abbreviations in NMR
spectroscopy
[133]
.


Figure 2.11
29
Si{
27
Al} REAPDOR pulse scheme. The
29
Si
27
Al dipolar coupling is re-introduced by the adiabatic
pulse on the
27
Al channel using a pulse length of 1/3 T
r
. The experiment employs
1
H decoupling during the dipolar
evolution and the detection periods in order to remove (or reduce) strong hetero-nuclear dipolar couplings between
the observed nuclei and the
1
H atoms present in the system.

As mentioned in Section (2.2.1), the assignment of
29
Si resonances from standard single-
pulse
29
Si MAS NMR may be somewhat uncertain due to the overlap of chemical shifts for
different SiO
4
environments caused by the Si
4+
Al
3+
substitution in their second-coordination
sphere. In this context, the REAPDOR experiment may be an appropriate tool to distinguish
Chapter 2. Applications of solid-state NMR in cement research 44
between different SiOAl connectivities and thereby, provides new insight to the interpretation
of the
29
Si MAS NMR spectrum
[134,135]
. From the general structural features of
aluminosilicates
[136]
, it is reasonable to consider an average SiAl internuclear distance (r
Si-Al
)
for SiOAl bonds in the Q
4
network structures since the bond lengths and bond angles of SiO
4

and AlO
4
tetrahedra only vary slightly for different SiOAl connectivities. This assumption
reduces equation (2.23) to
( )
2
r
0
nT =
Al
n k
S
S
(2.25)
where k represents all the constant parameters, including r
Al-Si
, and n
Al
is the number of Al atoms
substituted in the second coordination sphere of the observed silicon site.
The applicability of the
29
Si{
27
Al} REAPDOR experiment is tested for a synthetic sample
of NaY zeolite. The
29
Si MAS NMR spectrum of the aluminosilicate network structure of the
zeolite shows four well-separated resonances in the spectral region from 80 ppm to 105 ppm
(Figure 2.12). The difference
29
Si{
27
Al} REAPDOR spectrum clearly demonstrates that the
resonance at o(
29
Si) = 100.0 ppm is unaffected by the reintroduction of the
29
Si
27
Al dipolar
coupling and therefore, it is assigned to a Q
4
(0Al) site. The remaining resonances exhibit
REAPDOR curves with distinct slopes, which are illustrated in Figure 2.13. The values of kn
Al

for each of the resonances were obtained from a fit of the REAPDOR fractions, AS/S
0
s 0.3, to
equation (2.24). The resulting values are kn
Al
= 0.16, 0.30 and 0.44 for the resonances at 94.8
ppm, 89.2 ppm and 83.9 ppm, respectively. However, it should be noted that it was necessary
to include a data point for AS/S
0
> 0.3 in the curve fit for the resonance at 83.9 ppm since it has
only one data point that fulfils the condition AS/S
0
0.3. The corresponding number of Al atoms
in the second-coordination sphere for each silicon site (i.e. n
Al
= 1, 2 and 3) are obtained by
taking the relative ratios of their kn
Al
values, since k is approximately constant for all Q
i
(nAl)
components in the network structure. Thus, according to their isotropic chemical shift values in
conjunction with the number of Al in their second-coordination sphere, the resonances at 94.8
ppm, 89.2 ppm and 83.9 ppm are unambiguously assigned to the Q
4
(1Al), Q
4
(2Al) and
Q
4
(3Al) sites, respectively. This assignment is consistent with the structure of NaY zeolite
proposed from earlier studies
[80,137]
.

Chapter 2. Applications of solid-state NMR in cement research 45

Figure 2.12
29
Si{
27
Al} REAPDOR spectra (14.1 T, v
R
= 10.0 kHz, nT
r
= 1.2 ms) for a synthetic sample of NaY
zeolite: (a) reference spectrum (S
0
), (b) REAPDOR spectrum (S) in which the
29
Si
27
Al dipolar couplings have been
reintroduced and (c) the difference spectrum (S
0
S).



Figure 2.13
29
Si{
27
Al} REAPDOR curves for a synthetic sample of NaY zeolite. The experimental REAPDOR
fractions AS/S
0
as a function of evolution times (T
r
= 0.1 ms) are shown for the resonances at () 83.9 ppm, ()
89.2 ppm, (-) 94.8 ppm and () 100.0 ppm. The curve fits of the REAPDOR fractions for small dephasing
(AS/S
0
s 0.3) using the function AS/S
0
= kn
Al
(nT
r
)
2
are shown as solid lines.
Chapter 2. Applications of solid-state NMR in cement research 46
2.2.10. Multiple-Quantum (MQ) MAS NMR
The MQMAS NMR experiment
[138]
plays a very important role in several recent
investigations of half-integer spin quadrupolar nuclei. In general, the analysis of NMR spectra
for the central transition of such spin nuclei is complicated by a strong second-order quadrupolar
interaction. This interaction may introduce a severe line-broadening of the observed signals,
causing an overlap of resonances from structurally different sites. On the other hand, in a two-
dimensional (2D) MQMAS NMR experiment, the second-order quadrupolar line-broadening is
effectively removed in the isotropic dimension (F1) whereas it is retained in the anisotropic
dimension (F2). The elimination of the second-order quadrupolar broadening in the MQMAS
experiment significantly enhances the NMR spectral resolution for quadrupolar nuclei, enabling
more complex quadrupolar spin systems to be studied. Several modifications of the MQMAS
experiment have been developed with the aim of improving the sensitivity of the experiment
[139-
143]
. It has also been applied with success in combination with other rf pulse schemes such as in
the CP-MQMAS to obtain a high-resolution hetero-nuclear correlation (HETCOR) spectrum
[144]
.
This section gives a brief description of the basic two-pulse sequence for the triple-quantum
MQMAS experiment (Figure 2.14). However, the concepts may be applied to any other multi-
quantum excitation schemes.
The triple-quantum MQMAS experiment utilizes the fact that both the single- (1/2
1/2) and the triple- (3/2 3/2) quantum transitions are affected by the second-order
quadrupolar interaction but unaffected by the strong first-order quadrupolar interaction. In the
MAS axis system, the time averaged precession frequency due to the quadrupolar interaction for
such symmetric transitions (m m) is governed by
[145]


( ) ( ) ( ) ( ) ( )
( ) ( ) ( ) cos ,
cos , , ,
4
(4)
I
(4)
Q
2
(2)
I
(2)
Q
(0)
I
(0)
Q Q
P
P
m C
m C m C m I
+
+ =
(2.26)
( ) ( ) 1 3cos
2
1
cos
2
2
= P
( ) ( ) 3 30cos 35cos
8
1
cos
2 4
4
+ = P
where u is the angle between the spinning axis and B
,
and C
(i)
I
are constant factors which
depend of m and the spin-quantum number I. o and | are the Euler angles describing the
orientation of the quadrupolar PAS relative to the rotor-axis MAS system. P
2
(cosu) and
Chapter 2. Applications of solid-state NMR in cement research 47
P
4
(cosu) are the second-order and fourth-order Legndre polynomials. It can be seen from
equation (2.26) that the geometrical dependency (3cos
2
u 1) of the second-order Legndre
function can be removed by applying 2 arctan = , i.e. the magic angle. Furthermore, the term
depending on the fourth-order Legndre polynomial may be eliminated by correlation of two
different symmetric transitions that fulfils the condition
0
2 2
4
I 1 1
4
I
= + )t (m C )t (m C
) ( ) (
(2.27)
In practice, the MQMAS NMR experiment employs an increment of t
1
and the signal is acquired
as a function of t
2
. Therefore, m
2
must be 1/2 whereas m
1
can be freely selected by applying the
rules of phases cycling. Figure 2.14 shows the basic MQMAS pulse sequence for triple-quantum
excitation (i.e. m
1
= 3) with the corresponding coherence transfer pathway for a half-integer
spin quadrupolar nucleus. It can be seen from equation (2.27) that the spin evolution as triple-
quantum coherence under the effect of the second-order quadrupolar interaction for a time t
1
is
refocused after a time t
2
= t
1
k (where k = 7/9 and 19/12 for I = 3/2 and 5/2, respectively) after
the conversion of the triple-quantum coherence to single-quantum coherence.


Figure 2.14 Schematic illustration of the pulse sequence and the coherence transfer pathway for a triple-quantum
MQMAS NMR experiment. The selection of the triple- and single-quantum coherences are obtained by the
excitation and mixing pulses, which are phase-cycled in accordance with A|
i
= 2t/2N, where | is the phase of the
pulse and N is a multiple of Ap. The triple-quantum coherences excited by the first pulse are evolved for a time t
1

before these transitions are converted to single-quantum coherence by the second pulse. The acquisition begins
immediately after the second pulse to gain the best S/N ratio, although the echo occurs at a time t
2
= t
1
k (where k =
7/9 and 19/12 for I = 3/2 and 5/2, respectively) after the second pulse. Furthermore, the phase sensitivity in the
indirect dimension F1 is conducted by the hyper-complex phase-cycling method
[139]
.

Chapter 2. Applications of solid-state NMR in cement research 48
The advantage of the MQMAS NMR experiment is illustrated for a sample of gibbsite, -
Al(OH)
3
(Figure 2.15). The projection in the F2 dimension, under ideal conditions, corresponds
to a standard single-pulse
27
Al MAS NMR spectrum, in which the second-order quadrupolar
line-broadening has not been removed. The interpretation of this spectrum can be somewhat
uncertain since the observed second-order lineshape can potentially be misinterpreted as an
overlap of several resonances. On the other hand, the projection of the F1 dimension in
conjunction with the contour plot clearly identify two well-separated resonances, where the
shoulders at 4 ppm, 3.5 ppm and 9 ppm seen in the F2 dimension correspond to the
singularities and shoulders for the Al site with the strongest quadrupolar coupling. The
observation of two distinct crystallographic Al sites by MQMAS NMR is consistent with the
crystal structure for gibbsite determined by XRD
[146,147]
.


Figure 2.15
27
Al MQMAS NMR spectrum (9.4 T, v
R
= 10.0 kHz) of gibbsite, -Al(OH)
3
. The spectrum was
recorded using the three-pulse z-filter sequence and a 144-steps phase cycle. Furthermore,
1
H decoupling (TPPM)
was employed during both the evolution of the triple-quantum coherence (t
1
) and the detection (t
2
) periods. The
asterisks (*) indicate spinning side bands in the F1 dimension. Projections onto F1 and F2 dimensions,
corresponding to summations, are shown on the left side and above the 2D spectrum, respectively.
Chapter 3. Fluoride Mineralization 49



3. Chapter
Fluoride Mineralization




Since its first application in clinker preparation, probably in the late 1800s by Michaelis,
fluoride mineralization has got much attention in cement research
[41,42,148]
. Due to its high natural
abundance, CaF
2
(e.g., fluorspar) is by far the most widely used fluoride mineralizing agent in
modern cement production. However, other fluoride-containing compounds such as NaF
2
, MgF
2
,
Na
2
SiF
6
and MgSiF
6
have also shown similar properties
[25,149]
. In addition to the mineralizing
properties, fluoride also exhibits fluxing properties and seems to be involved in several chemical
reactions during clinker formation. At a kiln temperature of about 1100 C, fluoride reacts with
the raw meal components forming a clinker melt
[150,151]
, largely consisting of
11CaO7Al
2
O
3
CaF
2
. As belite and lime are brought together within the liquid phase, alite is
slowly formed. The alite formation accelerates rapidly as the kiln temperature is raised just
above 1200 C; approximately 40 % of alite is already formed at 1200 C and about 60 % at
1300 C. At kiln temperatures above 1300 C, 11CaO7Al
2
O
3
CaF
2
decomposes to tricalcium
aluminate and CaF
2
. It has been demonstrated in earlier studies of synthetic alite that a small
amount of fluoride can be incorporated in this phase, most likely in accordance with an O
2

F

substitution. Charge balance can be preserved by the incorporation of a similar amount of


Al
3+
ions substituting for Si
4+
in the alite structure, tentatively forming a solid solution with the
chemical formula Ca
3
[Si
1-x
Al
x
][O
5-x
F
x
], where the upper limit for x is approximately 0.15
[9]
.
This chapter presents the results from a study of the mineralizing effect of fluoride on the
formation of the calcium silicate phases of Portland cement. Furthermore, the site preference of
F

and its coupled substitution mechanism with Al


3+
and Fe
3+
ions have been investigated using
solid-state NMR. Finally, the results are applied in an optimization of the alite content in
Portland clinker.
Chapter 3. Fluoride Mineralization 50
3.1. Site preferences of F

ions in the calcium silicate phases of Portland cement


3.1.1.
19
F MAS NMR
The
19
F MAS NMR spectrum of a commercial white Portland clinker (wPc) containing
only 0.04 %w F is shown in Figure 3.1. The spectrum exhibits a characteristic lineshape
covering a spectral region from about 100 ppm to 125 ppm, which can only be satisfactorily
simulated by including at least three
19
F resonances with o(
19
F) = 122.1 ppm, 116.5 ppm and
111.4 ppm (Figure 3.1 b-c).
19
F MAS NMR spectra of modified clinkers prepared from wPc,
but containing higher F and Al
2
O
3
contents, show somewhat similar lineshape features with a
centre of gravity at o(
19
F) ~ 114.9 ppm, cf. Figure 3.10. This indicates that all F atoms occur in
similar structural environments, but in a less-ordered arrangement. Since the values of o(
19
F) for
fluoride ions cannot be distinguished from those for SiF and AlF covalent bonds, due to the
strong effect of different types of counter ions on the
19
F chemical shiftss, an exact identification
of the fluorine environments in Portland clinker cannot be extracted from the
19
F chemical shift
alone. However, their chemical shifts exclude them from being considered as the crystalline
phases Ca
4
Si
2
O
7
F
2
(o(
19
F) = 101.6 ppm and 106.1 ppm), CaF
2
(o(
19
F) = 105.9 ppm) or the
alumina-rich phase 11CaO7Al
2
O
3
CaF
2
, which exhibits two different
19
F resonances with a
large spinning sideband intensities resulting from the CSA interactions. Furthermore, recent
studies of aluminosilicate glasses demonstrated that AlF and SiF covalent bonds are only
found in glasses with much higher F contents as compared to that in Portland clinker
[152,153]
.

Figure 3.1
19
F MAS NMR spectrum (7.05 T, v
R
= 10.0 kHz) of a commercial white Portland clinker (wPc)
containing 0.04 %w F. The experimental spectrum (a) can be simulated, shown in (b) and (c), by three resonances at
o(
19
F) = 122.1 ppm, 116.5 ppm and 111.4 ppm.
Chapter 3. Fluoride Mineralization 51
3.1.2.
29
Si{
19
F} and
27
Al{
19
F} CP/MAS NMR
The incorporation of fluoride ions in the calcium silicate phases of Portland clinker has
been investigated for a series of fluoride-mineralized clinkers using
29
Si{
19
F} and
27
Al{
19
F}
CP/MAS NMR. Since the magnetization is transferred through the dipolar coupling (d 1/r
IS
3
),
the CP period acts as a filter which only allows the
29
Si and
27
Al spins that are located in the near
vicinity to a
19
F spin to be detected. The studied clinkers were modified from a commercial
white Portland clinker to have a high Al
2
O
3
bulk content of 4.3 %w, a lime saturation factor
(LSF) of 0.90 and fluoride contents in the range of 0.04 1.0 wt. % F. The bulk fluoride
contents in selected samples have been measured with an ion-selective electrode, after the
samples was dissolved in a mixture of HCl and KAl(SO
4
)
2
. These clinkers are used as reference
samples in the quantification of the fluoride content from
19
F MAS NMR experiments for the
remaining samples. Furthermore, the free lime content in all modified clinkers has been
measured. The experimental details for sample preparations and analyses are given in Appendix
1 and 2, respectively.


Figure 3.2
29
Si{
19
F} CP/MAS NMR spectra (7.1 T) of three selected fluoride-mineralized clinkers containing 0.23
%w F (a), 0.47 %w F (b) and 0.77 %w F (c). The spectra were recorded with a spinning speed of v
R
= 3.0 kHz and a
2.0 ms CP contact time. The overall lineshape observed in these spectra resembles closely the
29
Si MAS NMR
spectrum of alite in its monoclinic M
III
form, see Figure 2.
Chapter 3. Fluoride Mineralization 52
29
Si{
19
F} CP/MAS NMR spectra of selected modified clinkers are shown in Figure 3.2.
The absence of the belite resonance at o(
29
Si) = 71.33 ppm in these spectra demonstrates that
the fluoride ions are only incorporated in the alite phase. Furthermore, they all exhibit broadened
lineshapes with features resembling closely that of alite in the M
III
form. In contrast to this work,
most of the previous studies on the structure of fluoride-mineralized alite propose that this phase
is stabilized in its rhombohedral form
[9,154,155]
. However, the polymorphic form of the alite phase
when present in real Portland cement may as well be controlled by several other factors
[26]
such
as the presence of other impurities, the cooling rate, etc.
As considered earlier, the O
2
F

substitution should be charge-balanced by a similar


amount of Al
3+
guest ions substituting into the tetrahedral silicon sites of alite. The location of
these Al
3+
guest ions are, of course, important for the thermodynamic properties of the fluoride
mineralization since a random distribution of F

and Al
3+
guest ions in alite yields the highest
possible number of atomic configurations and thereby, maximizes the entropy of mixing for this
phase. However, a random distribution of F

and Al
3+
guest ions will create local regions with
formula units (Ca
3
SiO
4
F)
+
and (Ca
3
AlO
5
)

, which are probably energetically unfavorable. On the


other hand, the charge balance can be preserved locally if Al
3+
guest ions are located in the
proximity to the fluoride ions. Consequently, the incorporation of these Al
3+
guest ions will not
affect the entropy of mixing for the alite phase significantly.
27
Al MAS and
27
Al{
19
F} CP/MAS
NMR spectra of a modified clinker containing 0.77 %w F are shown in Figure 3.3. The single-
pulse spectrum (Figure 3.3) shows a sharp
27
Al resonance at approximately 75 ppm, which
originates from the tetrahedrally coordinated Al
3+
guest ions in the calcium silicate phases. In
addition to the fact that fluoride is only incorporated in the alite structure, the appearance of the
resonance at o(
27
Al) ~75 ppm in the
27
Al{
19
F} CP/MAS spectrum (Figure 3.3 b), reveals that the
fluoride ions are located in the near vicinity of the Al
3+
guest ions. Furthermore, the HH-
matching condition of v
F
= 3v
Al
+ v
R
applied in this experiment to achieve magnetization
transfer reflects that the observed
27
Al spins possess a large quadrupole coupling constant, which
is consistent with the values reported previously for Al
3+
guest ions in the alite phase. The
observation clearly clarifies the essential role of an aluminum source in fluoride mineralization
as proposed by Shame and Glasser
[9]
, since charge balance is achieved locally by a coupled
incorporation of F

and Al
3+
ions.

Chapter 3. Fluoride Mineralization 53

Figure 3.3 (a)
27
Al MAS and (b)
27
Al{
19
F} CP/MAS NMR spectra (7.1 T and v
R
= 5.0 kHz) of a fluoride-
mineralized clinker containing 0.77 %w F. The magnetization transfer from the
19
F spins to the
27
Al spins is
obtained using a Hartmann-Hahn matching condition of v
F
= 3v
Al
+ v
R
and a contact time of 1.5 ms. The asterisks
(*) indicate spinning sidebands.

3.1.3.
29
Si{
19
F} CP-REDOR NMR
The structures of the alite polymorphs include two different crystallographic oxygen sites,
as described in section 1.2.1: the covalently bonded oxygen in SiO
4
tetrahedra (SiO
b
) and the
interstitial oxygen ions (SiO
i
). Their corresponding average SiO internuclear distances (r
Si-O
)
are 1.63 and 4.32 , respectively, for the monoclinic M
III
structure. The mineralizing properties
of fluoride ions may strongly depend on which types of those oxygen sites that are accessible for
O
2
F

substitution. In order to elucidate the site preference of fluoride ions in the alite
structure, the average internuclear distance of the SiF spin pairs has been measured using
29
Si{
19
F} CP-REDOR NMR. The initial CP period acts as a
29
Si
19
F filter by only detecting the
29
Si nuclear spins which are dipolar coupled to
19
F, i.e., roughly r
SiF
< 5 . Subsequently, the
dipolar re-coupling ability of the REDOR sequence allow a measurement of the
29
Si
19
F dipolar
coupling constant, from which the average SiF internuclear distance can be calculated.
Furthermore, the CP period enhances the sensitivity of the experiment since the magnetization is
initiated from the
19
F spins, which have a high gyromagnetic ratio and a short relaxation delay of
only 8 s.
Chapter 3. Fluoride Mineralization 54

Figure 3.4
29
Si magnetization (M
x
(t) given in arbitrary units) as a function of the CP contact time obtained from
29
Si{
19
F} CP/MAS NMR experiments (7.1 T, v
R
= 5.0 kHz) for a modified clinker containing 0.77 wt.% F () and
for cuspidine (+). The experiments employed rf fields of
Si
B
1
/2t = 47.5 kHz and
F
B
2
/2t = 42.1 kHz, and a
relaxation delay of 8 s. The solid curves illustrate the optimum fits to the experimental data for the clinker and
cuspidine to equation (3.1) and (3.2), respectively.

The cross-polarization dynamics curves for a modified clinker containing 0.77 %w F and
a synthetic sample of cuspidine are shown in Figure 3.4. Considering the cross-polarization
curves for the applied contact times (t s 8.0 ms), the cross-polarization time constant (T
SiF
),
which reflect the strength of the SiF dipolar couplings, may be obtained by fitting the
experimental data to
[106]

M
x
(t) = [M
0
/(1 T
SiF
/T
1,F
)| exp(t
cp
/T
1,F
) [1 exp((1/ T
1,F
1/T
SiF
) t
cp
)] (3.1)
and
M
x
(t) = M
0
[1 exp( t
cp
/T
SiF
)] (3.2)
for the clinker and cuspidine, respectively. In equation (3.1), it is assumed that the
29
Si rotating-
frame relaxation time is at least an order of magnitude larger than the corresponding value for
19
F (i.e., T
1,Si
>> T
1,F
). On the other hand, equation (3.2) assumes that both T
1,Si
and T
1,F
are
much larger than T
SiF
. M
0
is the
19
F magnetization immediately after the initial 90 pulse on the
19
F-channel and t
cp
is the CP contact time. An optimized fit corresponding to the parameters T
SiF

Chapter 3. Fluoride Mineralization 55
= 4.61 ms and T
1,F
= 4.57 ms has been obtained for the modified clinker. On the other hand, the
SiF
2
three-spin spin system of cuspidine, with SiF distances of 3.88 and 4.00 , exhibits a
CP time constant of only 1.20 ms. Owing to the low natural abundance of the
29
Si nuclear spins
(4.7 %) and the low fluoride content (0.77 %w) of the Portland clinker, it is reasonably to
consider the
29
Si
19
F spin-pairs as isolated. In the absence of spin diffusion, the SiF spin
system may be approximated by a rigid lattice model, where the cross-polarization time constant
T
SiF
becomes inversely proportional to the dipolar second moment M
2
, whose magnitude
depends on the SiF distance (
6
SiF
r ) and the number of F atoms involved in the dipole coupling
(see equation 2.24). According to this assumption, the factor of nearly four between the values of
the CP time constant for the SiF spin system in Portland clinker and the three-spin system,
SiF
2
, in cuspidine indicates that the r
SiF
distances in alite is slightly longer than that for SiF in
cuspidine structure.
Results from the
29
Si{
19
F} CP-REDOR experiment are summarized in Figure 3.5 and 3.6.
The experimental data are plotted along with their optimized simulation and two simulated
universal REDOR curves for the average SiO
b
and SiO
i
distances in monoclinic alite (Figure
3.6). As a result of the severe overlap of resonances and the low signal to noise ratio of the
REDOR spectra (Figure 3.5), it has not been possible to distinguish between the different
29
Si
resonances from the eighteen crystallographic silicon sites of monoclinic alite. Therefore, they
have been considered to be equal in the analysis of the REDOR spectra. A dipolar coupling
constant of d = 285 Hz, corresponding to an average internuclear distance of 4.29 for the
isolated SiF spin-pair in alite, was obtained from a numerical fitting of the experimental
REDOR fractions to equation (2.21), see Appendix 2. The magnitude of the SiF internuclear
distance as well as the slopes of the REDOR curves shown in Figure 3.6 clearly demonstrate that
the fluoride anions are located in the interstitial oxygen site, i.e., the fluoride anions not involved
in covalent SiF bonds. These data are obtained for an optimized CP contact time of 3.0 ms
(Figure 3.4). However, selected REDOR experiments for CP contact times of 1.0 and 7.0 ms
give AS/S
0
fractions of the same order of magnitude as observed for the 3.0 ms contact time.
The experimental observation for the fluoride environments in the alite phase of Portland
cement is consistent with a recent theoretical investigation of the coupled incorporation of F

and
Al
3+
guest ions in the alite phase applying Density Functional Theory (DFT) calculation and
structure optimizations
[156]
. These calculations of energies for the Si
4+
+ O
2
Al
3+
+ F

substitution propose that the F

ion is preferentially substituted for the interstitial oxygen site,
since it requires a much lower substitution energy as compared to the value for the formation of
SiF covalent bonds.
Chapter 3. Fluoride Mineralization 56

Figure 3.5
29
Si{
19
F} CP-REDOR NMR spectra (7.1 T, v
R
= 10.0 kHz) of a fluoride-mineralized Portland cement
containing 0.77 %w F. (Left) Full signal (S
0
), which is not affected by the SiF dipolar coupling. (Right) The
attenuated signals (S), reflecting the intensity reduction caused by the SiF dipolar couplings. The evolution periods
are (a): 4 T
r
, (b): 16 T
r
, (c): 28 T
r
and (d) 40 T
r
, where T
r
= 0.1 ms.


Figure 3.6 Experimental REDOR fractions (AS/S
0
) for the
29
Si
19
F spin-pairs in the alite phase of a modified
Portland clinker (0.77 %w F). The experiments employ CP contact times of 1.0 ms (o), 3.0 ms (-) and 7.0 ms (x).
Numerical fitting of the experimental data (solid line) results in a r
Si-F
distance of 4.29 . The numerical simulations
for r
Si-F
of 1.6 and 4.3 are illustrated by the dashed lines, corresponding to the mean SiO bond length of the
SiO
4
tetrahedra and the mean distance for the interstitial oxygen atoms (O
i
) to their nearest Si atoms, respectively.
Chapter 3. Fluoride Mineralization 57
3.2. Mineralizing effects of calcium fluoride and calcium sulphate
The main effect of fluoride mineralization may be addressed to the fact that the coupled
incorporation of F

and Al
3+
ions increases the entropy of mixing ( ) ln( k
mix
S = ) for the alite
phase. However, due to the local charge-preserving effect of F

and Al
3+
, only the concentration
of F

contributes effectively to the increase in mixing entropy, where the total number of
possible atomic configurations
[157]
is given by ) ! ! /( !
O F al interstiti
n n n = ; n
interstitial
is the total
number of the interstitial sites, i.e. nine in the triclinic unit cell, Ca
27
Si
9
(O
b
)
36
(O
i
)
9
, while n
F
and
n
O
are the numbers of the interstitial sites occupied by F

and O
2
, respectively. Thus, the
entropy of mixing associated with fluoride mineralization for the alite phase has its maximum
when n
F
= 4.5, corresponding to the incorporation of 4.1 %w F in alite, or x = 0.1 for the formula
Ca
3
[Si
1-x
Al
x
][O
5-x
F
x
]. However, this content is much lower as compared to the fluoride level of x
= 0.15, reported in a previous study using the selective chemical dissolution method
[9]
. One
possible explanation for this discrepancy is that at high fluoride contents, the F atoms may also
form AlF and/or SiF covalent bonds. At sufficiently low fluoride levels, the ethalpy (H) for
the alite phase may not be affected by the presence of fluoride ions. On the other hand, the Gibbs
free energy (AG) for alite formation from belite and lime will be more negative
[157]

T = S H G (3.3)
( ) S H G , ,
lime belite alite
= I I + I I = I ,
where T is the temperature at which the reaction takes place. Alternatively, for complete reaction
between belite and lime, the amount of uncombined CaO, i.e. the free lime content, can be used
to monitor the thermodynamic stability of alite relative to belite. For a given burning condition, a
high free lime content indicates that belite is more favorable as compared to alite whereas a low
free lime content indicates the reverse.
Calcium sulphate is another widely use mineralizing agent in Portland clinker
production
[42,148,158]
, either alone or in combination with CaF
2
. In contrast to F

, the S
6+
ions are
preferentially incorporated in the belite structure
[38]
, in accordance with the tentative coupled
substitution mechanism, 3Si
4+
S
6+
+ 2Al
3+
. The mineralizing effect of CaSO
4
has been
reexamined for a series of modified clinkers prepared from raw meals including 4.3 %w Al
2
O
3

(~12 %w tricalcium aluminate), a LSF of 1.0 and SO
3
contents in the range 0 3 %w. The
29
Si
MAS NMR spectra for these samples are shown in Figure 3.7. It is apparent from the spectra that
the belite resonance exhibits an increasing line-broadening as a function of the SO
3
content,
Chapter 3. Fluoride Mineralization 58
supporting the assumption of an increased incorporation of S
6+
in the belite structure.
Furthermore, in addition to an increase in the belite intensity with increasing SO
3
content (Figure
3.7), the amount of uncombined CaO is increased rapidly as a function of the SO
3
content,
Figure 3.8. This clearly demonstrates that the presence of a SO
3
source in the raw meal increases
the stability of the belite phase and thereby, reduces the quantity of alite. It is also apparent from
Figure 3.8 that SO
3
has a rather high volatility since the actual SO
3
contents are much lower than
the SO
3
amount added to the corresponding raw meals.
In a similar way, the mineralizing effect of calcium fluoride is demonstrated for a series
of modified clinkers prepared from raw meals including a LSF of 1.0, 4.3 %w Al
2
O
3
, 3.0 %w
SO
3
(the actual SO
3
content in the clinkers is approximately 2.3 %w) and fluoride contents in the
range 0.04 1.0 %w F. The strong mineralizing effect of calcium fluoride is recognized by the
rapid decrease in the amount of uncombined CaO as a function of the fluoride content, Figure
3.8. It can be seen that the stability of SO
3
-mineralized belite is effectively reduced by the
addition of calcium fluoride. It only requires 0.6 %w F to retain the free CaO content at ~0.4
%w, i.e. the same level of uncombined CaO as in the clinker without using any mineralizing
agents. The mineralizing effects of fluoride and SO
3
on alite and belite, respectively, are
graphically illustrated in Figure 3.9.


Figure 3.7
29
Si MAS NMR spectra (7.1 T, v
R
= 7.0 kHz) of SO
3
-mineralized clinkers prepared from a commercial
white Portland clinker. The actual SO
3
content in the clinkers is 0.48 %w SO
3
(a), 1.09 %w SO
3
(b) and 1.91 %w
SO
3
(c). Furthermore, the preparation of the clinkers used a LSF of ~1.0 and a bulk Al
2
O
3
content of ~4.3 %w.

Chapter 3. Fluoride Mineralization 59

Figure 3.8 Mineralizing effects of CaSO
4
and CaF
2
reflected by the quantity of uncombined CaO (free lime) in the
clinkers as a function of the SO
3
and F contents, respectively. (Left) The graph shows the free CaO content () and
the actual SO
3
content () in the mineralized clinkers (LSF ~1.0, 0.04 %w F) as a function of the SO
3
quantity added
to the raw meal. (Right) Free lime content as a function of the actual fluoride content in the mineralized clinkers
(LSF ~1.0, 2.3 %w SO
3
). The actual SO
3
contents of the clinkers have been determined by X-ray fluorescence.


Figure 3.9 Graphical illustration of the effects of fluoride and SO
3
mineralization on the thermodynamic stability of
alite and belite, respectively.
Chapter 3. Fluoride Mineralization 60
3.3. Secondary effects of fluoride mineralization
In order to explore non-thermodynamic effects of fluoride mineralization, a series of
clinkers prepared from raw meals containing 4.3 Al
2
O
3
(~12 %w tricalcium aluminate), a LSF of
0.9 and fluoride contents in the range 0.04 1.0 %w have been investigated. Experimental
details including the sample preparation and analysis are given in Appendices 1 and 2.

The
19
F MAS NMR spectra of the clinkers show a single broad resonance with a centre of
gravity at 114.9 ppm, Figure 3.10. However, the lineshape exhibits some variation as a result of
the increased fluoride content. This observation confirms that for low concentrations (> 0.77
%w) the fluoride ions are only incorporated in the interstitial oxygen sites of alite, which is
consistent with earlier discussed investigations
[10,150]
, since excess phases of CaF
2
(o(
19
F) =
105.9 ppm), cuspidine (o(
19
F) = 101.6 ppm and 106.1 ppm) or 7CaO11Al
2
O
3
CaF
2
(two
19
F
sites with large CSAs) are not formed in these systems. However, due to the moderate volatility
of CaF
2
, the actual fluoride contents in the clinkers are somewhat lower than that used in the
clinker preparation.


Figure 3.10
19
F MAS NMR spectra (7.1 T, v
R
= 10 kHz) of selected fluoride-mineralized clinkers modified from a
commercial white Portland clinker to have a high bulk aluminate content of 4.3 %w (~12 %w tricalcium aluminate)
and fluoride contents of 0.04 (a), 0.23 (b), 0.36 (c) and 0.77 %w F (d). The broad resonance with a centre of gravity
at o(
19
F) ~ 114.9 ppm indicates that the fluoride anions have nearly identical structural environments, but they are
present in a less-ordered arrangement.
Chapter 3. Fluoride Mineralization 61

Figure 3.11
27
Al MAS NMR spectra (14.1 T, v
R
= 13.0 kHz) of selected fluoride-mineralized clinkers modified
from a commercial white Portland clinker (0.04 %w F, 2.13 %w Al
2
O
3
) to have a relatively high bulk aluminate
content of 4.3 %w (~12 %w tricalcium aluminate) and fluoride contents of 0.04 (a), 0.32 (b), 0.48 (c), 0.69 (d) and
0.77 %w F (e). The asterisks (
*
) indicate the spinning sidebands for spectrum (a) while the diamond (+) marks the
27
Al centerband from an AFt phase (ettringite), which most likely reflects a small degree of pre-hydration.

The correlation between the incorporation of fluoride and aluminum ions in alite is
apparent from the
27
Al MAS NMR spectra of the selected samples shown in Figure 3.11. The
increasing intensity for the
27
Al resonance at ~82 ppm illustrates that the increased incorporation
of fluoride ions in alite is accompanied by an increase in the quantity of Al
3+
guest ions in this
phase. On the other hand, the
27
Al resonance from the tricalcium aluminate phase, which appears
as a broad lineshape from approximately 30 ppm to 80 ppm, exhibits a decrease in its intensity
Chapter 3. Fluoride Mineralization 62
as the fluoride content is increased, indicating that this phase is partially consumed by the
increased incorporation of Al
3+
ions in alite. Therefore, the coupled substitution of F

and Al
3+

ions in alite may be described by the reactions
Ca
27
Si
9
(O
b
)
36
(O
i
)
9
+ 0.5xCaF
2
[Ca
27
Si
9
(O
b
)
36
(O
i
)
9-x
(F
i
)
x
]
x+
+ 0.5xCaO (3.4)
[Ca
27
Si
9
(O
b
)
36
(O
i
)
9-x
(F
i
)
x
]
x+
+ 0.5x3CaOAl
2
O
3

Ca
27
Si
9-x
Al
x
(O
b
)
36
(O
i
)
9-x
(F
i
)
x
+ xSiO
2
+ 1.5xCaO (3.5)
xSiO
2
+ 2xCaO x2CaOSiO
2
(3.6)
The reactions only consider the incorporation of F

and Al
3+
guest ions, assuming that no other
guest ions are present in the solid solution. The alite phase is expressed with the formula for one
unit cell of its triclinic polymorph
[20]
, in which it is discriminated between the covalently bonded
(SiO) and interstitial (non-bonded) oxygen sites. In fact, the presence of fluoride may also
results in incorporation of Al
3+
ions in the alite phase by other mechanisms since Figure 3.11d
indicates that the majority of the bulk Al
2
O
3
content of 4.4 %w (~ 8.610
-2
mol Al
3+
) is
incorporated in this phase. On the other hand, the fluoride content in this sample only
corresponds to 1.810
2
mol F

.
The distribution of Si atoms in the calcium silicate phases, quantified by the molar ratio
of Si in alite and belite, n
Si
[alite/belite], is achieved from the
29
Si MAS NMR spectra using a
deconvolution procedure described previously. Although the broad sub-spectra for M
I
and M
III

alite originate from an overlap of resonances from their eighteen different crystallographic
silicon sites, a satisfactory simulation can be achieved by only including nine different
29
Si
resonances, Figure 3.12. In this work, the M
I
lineshape has only been employed for the reference
clinker shown in Figure 3.12, which contains 0.04 %w F. The deconvolutions of the
29
Si MAS
NMR spectra for the remaining fluoride-mineralized clinkers include a sub-spectrum of alite in
its M
III
form, with small modifications only at the tails of the overall peak. The n
Si
[C
3
S/C
2
S]
molar ratio, resulting from the deconvolutions, as a function of the actual fluoride content is
illustrated in Figure 3.13. This graph clearly demonstrates that for a fixed chemical composition
an increased incorporation of fluoride ions in alite tends to decrease the alite to belite molar
ratio. However, this decrease in n
Si
[alite/belite] does not reflect a decrease in the actual alite
content since the quantity of uncombined CaO in these clinkers is rather constant, i.e. approx.
0.4 %w free CaO.
Chapter 3. Fluoride Mineralization 63

Figure 3.12 Simulations of
29
Si MAS NMR spectra for two modified clinkers with LSF ~0.9, 4.3 %w Al
2
O
3
and
0.04 %w F (left) and 0.32 %w F (right), prepared from a commercial white Portland clinker. (Left) The spectrum
simulation (b) includes one resonance at o(
29
Si) = 71.33 ppm for the belite phase (c) and a broad sub-spectrum of
alite in its M
I
form (d). (Right) For the fluoride-mineralized clinker (a), a sub-spectrum of the M
III
form of alite is
employed (d). Both simulations include only nine different
29
Si resonances for the sub-spectra of alite.
The results from the
19
F,
27
Al and
29
Si NMR experiments demonstrate first of all that the
incorporation of fluoride ions in the interstitial oxygen sites of alite increases the amount of Al
3+

guest ions substituting for Si
4+
on the tetrahedral sites, for which the tricalcium aluminate phase
acts as the aluminate source for the fluoride mineralization, equation (3.4) and (3.5). The
coupled substitution of F

and Al
3+
guest ions stabilizes local regions with the composition
Ca
27
Si
9-x
Al
x
(O
b
)
36
(O
i
)
9-x
(F
i
)
x
in the alite structure. Secondly, belite is formed from the formally
released quantities of SiO
2
and CaO, equation (3.6), according to the fact that free SiO
2

has not
been observed in the
29
Si MAS NMR experiments and that the free CaO content in the clinkers is
constant. Thus, the decreased n
Si
[alite/belite] molar ratio as a function of the fluoride content
may account for the Si
4+
Al
3+
substitution on the tetrahedral sites of alite and an increase in
the quantity of belite formed from the released SiO
2
at the same time. Moreover, equations (3.4
3.6) reveal that the alite content is almost unaffected by the quantity of fluoride whereas the
chemical composition of this phase is significantly changed by the fluoride mineralization. This
secondary effect may be compensated by additional CaO according to the fact that fluoride
mineralization increases the thermodynamic stability for alite only, Figure 3.8.
Chapter 3. Fluoride Mineralization 64

Figure 3.13 The molar ratio of Si atoms in the alite and belite phases, n
Si
[alite/belite], for the fluoride-mineralized
clinkers. The graph is based on data obtained from deconvolutions of the
29
Si MAS NMR spectra.
Chapter 3. Fluoride Mineralization 65
3.4. Incorporation of Fe
3+
ions in the calcium silicate phases of Portland cement
The iron content in ordinary Portland cement is typically around 3 %w Fe
2
O
3
, but much
lower in white Portland cement. The main constituent iron phase is ferrite (Ca
2
(Al
x
Fe
1-x
)
2
O
5
),
although a small amount of Fe
3+
may also be present as guest ions in tricalcium aluminate and
the calcium silicate phases
[29,31,159]
. Thus, these Fe
3+
guest ions may be involved in the fluoride
mineralizing process in a similar manner as the Al
3+
ions. Considering the
27
Al MAS NMR
spectrum of the modified white clinkers shown in Figure 3.11 (e), it can be seen that the majority
of Al
2
O
3
must be incorporated in the alite phase, partly associated with the fluoride
mineralization. Following this result, a series of fluoride-mineralized clinkers has been prepared
from raw meals containing 1.0 %w F and a LSF of 0.9. Furthermore, the bulk Al
2
O
3
content of
4.3 %w is partly replaced by additional Fe
2
O
3
, targeting Fe/Al molar ratios in the range of 0.05
0.8. The corresponding bulk Fe
2
O
3
contents are in the range of 0.3 4.0 %w. The potential of
Fe
3+
ions in replacing Al
3+
in the fluoride mineralizing process has been investigated for this
series of modified clinkers using
29
Si MAS NMR,
29
Si IR NMR and XRD. This study also
reveals the influences of an increased incorporation of Fe
3+
ions in the calcium silicate phases of
Portland cement, with the main findings presented below.


Figure 3.14
29
Si MAS NMR spectra (7.05 T, v
R
= 5.0 kHz) of two modified clinkers (LSF ~0.9, 0.9 %w F)
containing 0.38 %w Fe
2
O
3
(a) and 3.33 %w Fe
2
O
3
(b). The spectra demonstrate that the
29
Si resonances from belite
as well as alite are strongly affected by the increased bulk Fe
2
O
3
.
Chapter 3. Fluoride Mineralization 66

The
29
Si MAS NMR spectra of two modified clinkers with different bulk Fe
2
O
3
contents,
shown in Figure 3.14, clearly demonstrate that the presence of paramagnetic Fe
3+
ions in
Portland cement introduces a significant line-broadening and intensity loss on the observed
nuclear spins. Recently, it has been demonstrated that the spin-lattice relaxation process for the
29
Si spins in the calcium silicate phases of ordinary Portland cements is predominantly affected
by the Fe
3+
ions incorporated as guest ions in these phases
[78]
, rather than by the Fe
3+
ions present
in the separate ferrite phase as proposed earlier
[160]
. This observation been reexamined for two
modified Portland clinkers containing 4.3 %w Al
2
O
3
, LSF ~ 1.0 and Fe
2
O
3
contents of 1.5 %w
and 2.0 %w. The aluminate and ferrite phases were removed from the clinkers using a selective
dissolution method
[161]
. The validity of this method is supported by the absence of the
2u reflexions at ~33.2 and ~33.8 in the XRD pattern of the clinkers after selective dissolution,
Figure 3.15. On the other hand, the
29
Si IR NMR spectra shown in Figure 3.16 for the same
clinkers before and after selective dissolution show no significant differences, which
demonstrates that the ferrite and aluminate phases only have a small effect on the
29
Si relaxation.

0
2000
4000
6000
8000
10000
12000
14000
16000
28 29 30 31 32 33 34 35
2 u
A
r
b
i
t
r
a
r
y

U
n
i
t
C
3
S
C
3
A
C
4
A
F
C
3
S
C
3
S
C
3
S
C
3
S

Figure 3.15 XRD diffractograms of a modified clinker (4.3 %w Al
2
O
3
, ~2.0 %w Fe
2
O
3
and LSF 1.0). The XRD
pattern for the original clinker is shown by the solid line while the dashed line corresponds to the pattern after
selective dissolution of the aluminate (C
3
A) and ferrite (C
4
AF) phases.
Chapter 3. Fluoride Mineralization 67


Figure 3.16
29
Si IR NMR spectra (7.1 T, v
R
= 5.0 kHz) of a modified clinker (4.3 %w Al
2
O
3
, ~2.0 %w Fe
2
O
3
and
LSF 1.0). The
29
Si IR NMR spectra of the same sample before and after selective dissolution are shown in parts
(a) and (b), respectively, where the recovery times (in seconds) are shown below the individual spectra.

Figure 3.17 shows the results from the
29
Si IR NMR experiments, including experimental
data together with their best fits to equation (2.18), for the alite and belite phases of three
clinkers with bulk Fe
2
O
3
contents of 0.38, 0.93 and 3.91 %w. The M
z
(t) values for the alite and
belite phases have been obtained for each of the
29
Si IR NMR spectra employing the
deconvolution procedure described earlier. It shows clearly that the relaxation times for the
29
Si
Chapter 3. Fluoride Mineralization 68
resonances from alite as well as belite are affected by the increase in the bulk Fe
2
O
3
content,
reflecting an increased incorporation of the paramagnetic Fe
3+
ions in the alite and belite
structures. However, the actual concentration of Fe
3+
ions incorporated in the alite and belite
phases can not be obtained from these experiments since the spin-lattice relaxation time for the
electron spins has not been determined, cf. equation (2.19). It is evident from the plot of
29
Si
spin-lattice relaxation times (T
1
) for alite and belite as a function of the bulk Fe
2
O
3
content,
Figure 3.18, that the increase in the amount of Fe
3+
guest ions (1/T
1
N
P
2
) is very similar for
these two phases at low bulk Fe
2
O
3
concentrations (i.e. < 0.8 %w Fe
2
O
3
), indicating that the
presence of F

ions in alite has no significant impact on the Fe


3+
guest-ion incorporation in the
calcium silicate phases. Therefore, Fe
3+
seems not to be involved in the fluoride mineralization.
The slight change in slope for 1/T
1
for bulk Fe
2
O
3
contents above 0.8 %w suggests that different
amounts of Fe
3+
ions are incorporated in alite and belite, potentially in different sites as
mentioned by Taylor
[13]
who suggests that the Fe
3+
ions enter the tetrahedral sites in belite by
substituting for Si
4+
ions, but occupy the octahedrally coordinated sites in alite obtained by
substitution for Ca
2+
ions.



Figure 3.17 Plots of M
z
(t) values for alite (left) and belite (right) as a function of the recovery time obtained from
the
29
Si IR NMR experiments for three modified clinkers containing 0.38 %w Fe
2
O
3
(), 0.93 %w Fe
2
O
3
() and
3.91 %w Fe
2
O
3
(). The best fits of the experimental data to equation 2.13 are shown as solid lines.


Chapter 3. Fluoride Mineralization 69

Figure 3.18 A plot of the relaxation rate (1/T
1
) for the
29
Si spins in alite () and belite (+) versus the bulk Fe
2
O
3

content for a series of modified clinkers. The clinkers were prepared from a commercial white Portland cement,
targeting LSF ~0.9, 0.9 %w F and Fe/Al molar ratios in the range of 0.05 0.8. The relaxation time constant T
1
for
each clinker was obtained from
29
Si IR NMR experiments.

The proposed Ca
2+
Fe
3+
substitution is further supported by an evaluation of the alite and
belite content for each of the studied modified clinkers by a plot of the molar ratio of silicon in
alite and belite (n
Si
(alite/belite)) as a function of the bulk Fe
2
O
3
content, Figure 3.19. For small
concentrations (< 0.8 %w Fe
2
O
3
), the incorporation of Fe
3+
ions in the calcium silicate phases
results in an increase of the alite content, i.e. the n
Si
(alite/belite) ratio, for a fixed clinker
composition apart from the variation in the Fe/Al molar ratio. Since the amount of free CaO is
nearly constant (~0.4 %w) for all clinkers, the increase in the alite content may potentially reflect
that the Fe
3+
ions preferentially substitute for the Ca
2+
sites in alite, corresponding to an effective
increase in the bulk CaO content. For a further increase of the bulk Fe
2
O
3
content above ~0.8
%w, the alite content decreases as a result of the formation of the ferrite phase which reduces the
amount of CaO available for alite formation. This is supported by the observation of the ferrite
phase for the clinkers with bulk Fe
2
O
3
contents above 0.8 %w by the 2u reflection at 33.8 in the
XRD diffractograms (not shown).
Chapter 3. Fluoride Mineralization 70

Figure 3.19 Plot of the molar fractions of silicon in alite and belite (-), n
Si
(alite/belite), as a function of the bulk
Fe
2
O
3
content. The n
Si
(alite/belite) ratios were obtained for each of the modified clinkers by deconvolution of the
29
Si MAS NMR spectra.

Chapter 3. Fluoride Mineralization 71
3.5. Hydration of fluoride-mineralized Portland cement
In addition to the strong mineralizing effect, the quantity of fluoride ions also shows an
important impact on the hydrational properties of the mineralized cement. It seems to have an
adverse effect on the reactivity of the calcium silicates leading to a retarding effect on the
hydration
[154,162]
. However, in combination with an optimized amount of SO
3
containing
mineralizers (e.g. CaSO
4
), fluorine shows improved effects on the early hydration process
[10,41]

with an optimum fluoride content, as monitored by strength measurements after one day, at
about 0.2 %w for an ordinary Portland cement. Higher fluoride contents will reduce the
compressive strength measured after one day while the long-term hydration, i.e. > 28 days, is
almost unaffected.


Figure 3.20 Experimental (a) and simulated (b)
29
Si MAS NMR spectra of a fluoride-mineralized cement hydrated
for 28 days. The cement is prepared from a clinker containing 0.36 %w F, 4.3 %w Al
2
O
3
and with a LSF of 0.9. The
deconvolution of the spectrum includes the sub-spectrum for the M
III
alite and resonances for belite and the CSH
phase. Alite and belite are shown in (c) while the CSH spectrum (d) includes five
29
Si resonances at o(
29
Si) =
76.7 ppm, 79.1 ppm, 81.5 ppm, 83.3 ppm and 85.4 ppm.
Chapter 3. Fluoride Mineralization 72
To investigate the effects of fluoride ions on the hydration process, hydrated cement
samples have been prepared (water/solid = 0.5, n
SO3
/n
Al2O3
= 1.3 and a relative humidity of RH =
100 %) for clinkers produced from raw meals including 4.3 %w Al
2
O
3
, a LSF of 0.9 and fluoride
content in the range 0.04 1.0 %w. The hydration has been stopped and studied by
27
Al and
29
Si
MAS NMR for each series at 1, 2, 7, and 28 days. The degrees of hydration, given by equation
2.11, for alite and belite at each hydration time were estimated from
29
Si MAS NMR spectra
using the common deconvolution procedure. Based on the observations for synthetic CSH
presented in Chapter 4, five
29
Si resonances at o(
29
Si) = 76.7 ppm, 79.1 ppm, 81.5 ppm,
83.3 ppm and 85.4 ppm have been included in addition to the resonances from belite and alite
in its M
III
form in the deconvolution of
29
Si MAS NMR spectra the for hydrated samples. An
example is shown in Figure 3.20 for a modified cement (0.36 %w F) hydrated for 28 days.
A plot of the degree of hydration as a function of the fluoride content for alite and belite
is shown in Figure 3.20. This graph clearly demonstrates that the quantity of fluoride ions has an
important impact on the hydration rate of the calcium silicate phases. A critical fluoride content
seems to exist around 0.36 %w F. Below this value, the fluoride ions enhance the hydration rate
for the alite phase, Figure 3.21. An increase in the fluoride content above 0.36 %w F
significantly retards the early hydration rate of alite. However, the effects from an increased
fluoride content on the hydration of alite are less apparent as the hydration proceeds and nearly
cancel out at 28 days of hydration. The degree of hydration for belite at 28 days of hydration also
tends to be reduced slightly by fluoride levels above 0.36 %w, Figure 3.21.

27
Al MAS NMR spectra of selected samples, investigating the effects of fluoride ions
on the hydration process of the Al-containing phases, are shown in Figure 3.22. It is apparent
from these spectra that the quantity of fluoride ions strongly influences the hydration of the
aluminate phases after one day of hydration. However, the effect from fluoride anions is less
apparent as the hydration proceeds to seven day of hydration, Figure 3.21, and it nearly cancels
out after hydration for 28 days (not shown). Considering the centers of gravity for the
centerbands at o
cg
(
27
Al) ~ 13 and 9 ppm in Figure 3.21, it can be seen that the AFt phase
(ettringite) is rapidly converted into an AFm phase (monosulphate) as the fluoride content is
increased up to 0.36 %w. Finally, when considering the intensity of the
27
Al resonance at ~82
ppm (i.e., Al
3+
guest-ions in alite), the fluoride content of 0.36 %w F also appears to be a critical
limit for the reactivity of alite. Below this value, an increase in the fluoride content tends to
enhance the hydration rate of the alite phases.
Chapter 3. Fluoride Mineralization 73


Figure 3.21 The degree of hydration for alite and belite in fluoride-mineralized cements prepared from clinkers
containing different quantities of fluoride ions. The relative quantities of the silicate species are obtained from
29
Si
MAS NMR spectra (7.1 T, v
R
= 7.0 kHz, 30-s relaxation delay) of the hydrated samples. (Top) the total degree of
hydration for alite after 1 day (), 2 days (-), 7 days () and 28 days () of hydration. (Bottom) the degree of
hydration for belite () after 28 days of hydration. The lines are shown as a guide to the eyes.
Chapter 3. Fluoride Mineralization 74
The hydration experiments demonstrate that the quantity of fluoride ions in Portland
cement has an important impact on the early hydration, most significantly after one and two days
of hydration. In agreement with the previous investigation for the effects of fluoride ions on the
compressive strength of Portland cement
[10]
, the
27
Al and
29
Si MAS NMR experiments of the
hydrated samples have shown the existence of a critical fluoride content, although for this series
of clinkers of the critical fluoride content is of ~0.36 %w F. Below the critical value, an increase
in the fluoride content tends to improve the hydration properties of the hydrated Portland
cement. Most importantly, the increasing fluoride content up to 0.36 %w F increases the
reactivity of the alite phase. Hence, a larger amount of aluminate ions may be released from the
hydration of alite into the paste-liquid, which probably is the reason for the rapid conversion of
the AFt (ettringite) into a AFm phase (monosulphate) after one day of hydration, see Figure
3.22. Above the critical value of ~0.36 %w F, the fluoride ions reduce the degree of hydration
for alite up to 28 days, although the most pronounced reduction is observed after one and two
days of hydration. One possible reason for the existence of this critical fluoride content is that
calcium fluoride has a very low solubility product of k
sp
= 3.710
11
M
3
, which causes CaF
2
to
precipitate immediately on the cement grain surface as the fluoride ions are released from the
hydration of alite (cf. Chapter 4). Above the critical fluoride content, the precipitating CaF
2

forms a protective layer on the cement grain surface reducing the cement hydration. This is
consistent with the observation by
29
Si and
27
Al MAS NMR revealing that above the critical
concentration, fluoride ions reduce the degree of hydration of alite and belite, and they affect the
formation of all hydrated phases. Thus, the presence of fluoride ions in high concentrations may
result in a decrease in the quantity of CSH phases and therefore, the early strength
development for the cement is reduced. Furthermore, it also retains a large amount of aluminium
ions in the anhydrous alite phase leading to a retarding effect on the conversion of AFt
(ettringite) into AFm phases. The critical fluoride content may, however, vary slightly depending
on the quantity and reactivity of the alite phase. Another effect that also needs consideration is
the decreased SiO
2
content in fluoride-mineralized alite as a consequence of the coupled
incorporation of F

and Al
3+
ions (section 3.2). Obviously, this results in a reduction in the
quantity of the CSH phase formed during the early hydration period, since only a small
amount of belite has reacted with water at this time. As the hydration process proceeds, the
cement grain will expand due to the gradual hydration of alite and belite. Hence, the protecting
layer of CaF
2
is broken and the retarding effects from the fluoride ions become less apparent
with time.
Chapter 3. Fluoride Mineralization 75

Figure 3.22
27
Al MAS NMR spectra (14.1 T, v
R
= 13.0 kHz) of selected hydrated fluoride-mineralized cements
after one (left column) and seven (right column) days of hydration. The cements were prepared from clinkers
containing 0.04 %w F (a), 0.23 %w F (b), 0.36 %w F (c) and 0.65 %w F (d). The spectra demonstrate that the
quantity of fluoride significantly affects the formation of the AFt (ettringite) phase and its conversion into AFm
phases (monosulphate). The asterisks (*) indicate the spinning sidebands from the AFt phases.
Chapter 3. Fluoride Mineralization 76
3.6. Application of fluoride mineralization
3.6.1. Preparation of test clinkers
Based on the results presented in the previous sections, a fluoride-mineralized clinker,
denoted test clinker, was prepared using a commercial white Portland clinker (HKL) as the main
source. The chemical compositions of the original and modified clinkers are shown in Table 3.1.
As it is shown by the table, only the contents of SiO
2
, Al
2
O
3
, Fe
2
O
3
and F

have been modified.


The modification of the chemical composition leads to an increase in the lime saturation factor
from 0.95 for HKL to approximately 1.00, although the actual CaO content is nearly constant for
the clinkers. Furthermore, the actual fluoride content in the test clinker is 0.28 %w rather than
0.36 %w which is the optimum fluoride content observed in the hydration experiments (cf.
section 3.5). This is due to the fact that CaF
2
has a moderate volatility, which makes it difficult
to control the actual content of fluoride in the final clinker. Quantitative Rietveld analysis of the
test clinker for batch 1 reveals that it consists of 77 %w alite, 21 %w belite and 2 %w tricalcium
aluminate whereas the unmodified HKL contains 69 %w alite, 24 %w belite and 5 % tricalcium
aluminate. In fact, the alite content may be further increased by additional CaO, if it is desired;
the saturation condition defined by equation (1.1), i.e. LSF = 1.0, may not be applied to this
clinker type since the majority of aluminum is not present in tricalcium aluminate but rather as
guest ions in the alit phase.

Table 3.1 Chemical composition of the original and the modified clinkers from XRF experiments. The bulk fluoride
content and the quantity of free CaO for the clinkers were measured using the methods described in Appendix 2.
HKL Test clinker
Batch 1 Batch 2
CaO 70.3 70.4 70.3
SiO2 25.4 22.8 22.5
Al2O3 2.13 4.49 4.40
Fe2O3 0.37 0.78 0.73
F 0.04 0.28 0.29
SO3 0.12 0.36 0.40
MgO 0.63 0.56 0.73
Na2O 0.19 0.10 0.07
K2O 0.07 0.00 0.02
Free CaO 1.7 0.4 0.4
Chapter 3. Fluoride Mineralization 77
3.6.2. Strength performances of fluoride-mineralized Portland cement
The clinkers were milled together with gypsum to produce cements with a Blaine fineness
of 480 m
2
/kg. The SO
3
content for the clinkers was optimized by following the heat evolution of
the hydration process up to eighteen hours after mixing for a series of cements with different
SO
3
contents. For white Portland cement (wPc) produced from the unmodified HKL, the
optimum SO
3
content is 2.1 %w whereas the test cements required a SO
3
content of 4.2 %w to
avoid flash setting, due to their rather high bulk Al
2
O
3
content. Strength tests were performed for
the white Portland cement, an ordinary Portland cement (oPc) and the two test cements (from
batch 1 and 2) in accordance with the mini-RILEM method
[163]
. The compressive strengths for
the hydrated cements were measured at 1, 2, 28 and 90 days of hydration with the results
presented in Figure 3.23. A comparison of the hydration experiments for wPc and oPc clearly
demonstrates the effect of the quantities of alite and belite on the strength development. For one
and two days of hydration, the high alite content in oPc results in higher compressive strengths
as compared to wPc. On the other hand, as the hydration proceeds the strength development is
taken over by the hydration of belite and therefore, wPc exhibits higher strengths at 28-days and
later hydration times.
After one day of hydration, the test cements exhibit comparable strengths as the oPc due
to their high alite content. However, the high belite content of the test cements as compared to
the oPc results in a much faster strength development during the long-term hydration. Moreover,
in comparison with the unmodified wPc, the test cements have shown improved compressive
strengths for all hydration stages; after one day of hydration, the test cements exhibit an increase
in the compressive strength corresponding to an average value of 36 % and approximately 10 %
after long-term hydration.
Strength tests have also been conducted for a series of blended cements produced from
the test clinker of batch 1, employing a 30 %w replacement of the clinker by limestone filler
(CaCO
3
), heat-treated bentonite or glass particles. The heat-treated bentonite and glass are
developed in two other PhD co-projects of the FUTURECEM project. After one and two days of
hydration, the test cement shows the best performance with limestone filler alone, Figure 3.24.
However, its compressive strength is somewhat lower than that for pure wPc. At 28 days of
hydration and later, the partly replacement by 10 %w CaCO
3
and 20 %w glass results in
compressive strengths which are roughly 5 % lower than that of pure wPc. The replacement of
the clinker by heat-treated bentonite results in the lowest strength for all samples, except for that
after one-day of hydration.
Chapter 3. Fluoride Mineralization 78
0.0
20.0
40.0
60.0
80.0
100.0
120.0
1 2 28 90
days
C
o
m
p
r
e
s
s
i
v
e

s
t
r
e
n
g
t
h

(
M
P
a
)

Figure 3.23 Compressive strength of blended cements measured in accordance with the mini-RILEM method. The
test cements (, ) and white Portland cement () have a Blaine fineness of 480 m
2
/kg whereas it is 575 m
2
/kg for
the ordinary Portland cement oPc (). Thus, to be able to compare the results of test cements with oPc, the actual
compressive strength of oPc has been adjusted relating to the difference in their fineness
[18]
.

0.0
10.0
20.0
30.0
40.0
50.0
60.0
70.0
80.0
90.0
100.0
1 2 28 90
Days
C
o
m
p
r
e
s
s
i
v
e

s
t
r
e
n
g
t
h

(
M
P
a
)

Figure 3.24 Compressive strength of blended cements measured in accordance with the mini-RILEM method. The
cements were prepared with a clinker substitution level of 30 %w, i.e. () 30 %w limestone filler, () 10 %w
limestone filler and 20 %w pre-treated betonite and () 10 %w limestone filler and 20 %w glass particles. The
unmodified wPc is shown in red ().
Chapter 3. Fluoride Mineralization 79
3.7. Summary
Site preferences for F

and Al
3+
guest ions in the calcium silicate phases of Portland
clinker have been investigated using
19
F,
27
Al, and
29
Si MAS NMR combined with double-
resonance experiments such as
27
Al{
19
F} and
29
Si{
19
F} CP/MAS, and
29
Si{
19
F} CP-REDOR. It
is demonstrated that fluoride mineralization can be represented by the coupled substitution
Si
4+
+ O
2
Al
3+
+ F

in alite, where fluoride ions substitute for the oxygen on interstitial sites
and Al
3+
ions replace silicon on tetrahedral sites in the near vicinity of the F

ions. The coupled


substitution of F

and Al
3+
stabilizes local regions with the chemical composition
Ca
27
Si
9-x
Al
x
(O
b
)
36
(O
i
)
9-x
(F
i
)
x
, where the upper limit of x is 4.5.
A series of laboratory-synthesized fluoride-mineralized Portland clinkers have been
prepared and studied using solid-state NMR and other analytical tools such as XRD, XRF, free
lime measurement, etc. This work demonstrates that the use of fluoride-mineralizing agents in
clinker production promotes the formation of alite by two different mechanisms. The most
important effect is that the incorporation of fluoride in alite results in an increase in the entropy
of mixing, which corresponds to a reduction in the Gibbs free energy for this phase. Assuming
that the enthalpy of alite is unaffected by the incorporation of fluoride, the optimum fluoride
content in Portland cement is approx. 4.1 % by the weight of alite. Another important effect
arises from the mass balance of the coupled substitution Si
4+
+ O
2
Al
3+
+ F

in alite, which
results in an increase in the quantity of belite and thereby, further facilitates the forward reaction
CaO + belite alite.
The incorporation of Fe
3+
ions in calcium silicate phases of Portland clinker has been
investigated using
29
Si IR NMR. It is evident that small amounts of Fe
3+
ions are incorporated
into the alite as well as belite structures. However, the Fe
3+
ions do not appear to be involved in
fluoride mineralization since the presence of a high F

content in alite does not promote the


incorporation of Fe
3+
ions in this phase. This study also demonstrates that the Fe
3+
ions tend to
be incorporated on the octahedral sites of alite by substituting for the Ca
2+
ions.
The hydration experiments for fluoride-mineralized cements have proven the existence of
a critical bulk fluoride content of 0.36 %w. Below this value, an increase in the fluoride content
tends to improve the hydrational properties of Portland cement, e.g. it promotes the hydration of
alite and the conversion of AFt (ettringite) into AFm (monosulphate) phases. A further increase
in the fluoride content above this critical value significantly reduces the degree of hydration for
alite up to 28 days. Furthermore, it exhibits a retarding effect on the conversion of AFt into
Chapter 3. Fluoride Mineralization 80
AFm, although this is nearly cancelled out after seven days of hydration. The high fluoride level
also seems to affect the degree of hydration for the belite phase.
A test clinker has been prepared based on the results from the optimization of the fluoride
content in Portland clinker. Although the CaO content in the test clinker is comparable to the
unmodified white Portland clinker, i.e. 70 %w CaO, the test clinker has an increase in the alite
content by approximately 8 %w whereas the content of belite and tricalcium aluminate is
decreased with a similar amount.
Tests on the strength performances of fluoride-mineralized cements prepared from the
test clinker have been conducted in accordance with the mini-RILEM method. The test cements
exhibit an increase in its one-day compressive strength by 36 % as compared to the unmodified
white Portland cement. For longer hydration times, the test cement shows an increase of
approximately 10 %.
Blended cements with a 30% clinker replacement have been prepared based on the test
clinkers. Three systems have been studied: (i) 30 %w lime, (ii) 10 %w limestone and 20 %w
glass particles and (iii) 10 %w limestone and 20 %w pre-treated clay. The blended cement (i)
shows the best early strength performance whereas system (ii) has the best performance after
long-term hydration. In both cases, the blended cements exhibit a reduction of about 5 % as
compared to the compressive strength of the unmodified white Portland cement.
Chapter 4. Fluoride-ion environments in CSH 81









4. Chapter



Fluoride-ion environments in synthetic
CSH and hydrated Portland cement







The main aim of this work is to investigate the structural environments of fluoride guest
ions in the calcium silicate hydrate (CSH) phases and their influence on the hydration of
Portland cement. As it is demonstrated in a number of previous studies, in addition to the strong
mineralizing effects, fluoride ions also affect the hydration properties of Portland cement; they
lead to an extended setting time and a reduction in the early strength of hydrated Portland
cement
[10,41,154,162]
. Despite the obvious important effects of fluoride, the mechanism by which
fluoride ions influence the hydration of Portland cement is only poorly understood. Two main
proposals have been presented. The first considers the retarding effect as an intrinsic property of
fluoride-mineralized alite
[155]
, which is possibly stabilized in its rhombohedral form. This
polymorph has proven a prolonged dormant time but it also hydrates much faster than the
remaining alite forms. The second proposal explains the retarding effect of fluoride as a result of
the precipitation of CaF
2
salt on the cement-grain surface, preventing the cement to hydrate
[164]
.
Chapter 4. Fluoride-ion environments in CSH 82
This explanation seems to be more appropriate for the reduction in the degree of hydration of
Portland cement with fluoride contents above the critical value (cf. Section 3.5). However,
experimental identification of CaF
2
forming from the hydrating Portland cement has not yet been
possible, due to the very low fluoride content.
This chapter is divided into two parts. The first focuses on the incorporation of fluoride
and its influence on the structure of synthetic calcium silicate hydrates while the second part
presents the results from a study on the fluoride environments in hydrated Portland cement. The
observation contributes to a further understanding of the mechanism by which fluoride strongly
modifies the early hydration properties of Portland cement.


4.1. A brief description of the CSH structure from the T/J viewpoint.
Figure 4.1 shows a schematic two-dimensional representation of the principal layer of
CSH structure based on the tobermorite (T) structure
[62,66]
, consisting of a CaO layer
sandwiched between two sheets of dreierketten silicate chains. The upper part of the figure
represents an infinite silicate chain with periodicity of one bridging Q
2
B
and two Q
2
P
sites. If the
principal layers are not moved too far away from each other, such as in tobermorite 11-, the
Q
2
B
silicate units from two adjacent layers can inter-link forming Q
3
units
[165]
. Different types of
guest ions may be present in the CSH structure
[39,65,166,167]
. For example, the substitution of
Si
4+
by Al
3+
ions on the bridging sites is usually accompanied by incorporation of monovalent
alkali cations or Ca
2+
in the interlayer, preserving the charge balance. As it is shown in the lower
part of Figure 4.2, each Al
3+
incorporated in the bridging sites results in two Q
2
(1Al) units. The
principal layer consists of highly disordered finite silicate chains of length 3n 1, where n is an
integer whose value depends on the Ca/Si molar ratio and the degree of protonation (w/n). The
tetrahedral silicate units on the bridging sites and the defect sites are charge-balanced partly or
fully by the H atoms present in the interlayer region, forming the silanonl SiOH groups. For
high Ca/Si molar ratios, the bridging SiO
4
tetrahedra may also be replaced by calcium ions,
which are coordinated to seven oxygen atoms originating either from water molecules or
hydroxyl groups, Figure 4.1. The figure is only for the purpose of demonstrating the distinctions
between different SiO
4
connectivities and the distribution of OH

in the CSH structures; it is


not meant to accurately represent the actual arrangement of the atoms.
Chapter 4. Fluoride-ion environments in CSH 83


Figure 4.1 Schematic two-dimensional representations of the CSH structure derived from the tobermorite/jennite
(T/J) viewpoint. The upper part in (a) shows an infinite dreierketten silicate chain with SiO referring to as SiOSi
and SiO

groups that are charge-balanced by Ca


2+
present in the interlayer (not shown). The lower part of (a)
shows the incorporation of Al
3+
ions on the bridging site leading to two Q
2
(1Al) sites for each Al
3+
for Si
4+
substitution. The diagram shown in (b) illustrates the distribution of hydroxyl groups between tobermorite- and
jennite-based dimer structures. For the T-based CSH structure, OH

occurs in the interlayer (i.e. the lower part)


whereas it takes part of the principal layer (i.e. the upper part) for the jennite-based part.

Chapter 4. Fluoride-ion environments in CSH 84
4.2. Fluoride-ion environments in synthetic CSH and hydrated Portland cement
29
Si MAS and
29
Si{
19
F} CP/MAS NMR spectra of a fluoride-mineralized Portland
cement and two synthetic CSH samples, revealing the presence of F

guest ions in the CSH


phase, are shown in Figure 4.2 and 4.3, respectively. For the hydrated Portland cement sample
(Figure 4.2), the
29
Si resonances identifying SiF connectivities from remainders of alite appear
in the spectral region from about 65 ppm to 75 ppm while for the CSH phase, they are
observed from 75 ppm to 90 ppm
[82]
. The
29
Si{
19
F} CP/MAS NMR spectrum clearly shows
the presence of the Q
1
and Q
2
components. However, a clear identification of the remaining
silicate environments (such as Q
2
(1Al) and Q
2
B*
) cannot be extracted from the spectrum due to
the severe overlap of resonances from the disordered CSH structure formed by the hydration
of Portland.


Figure 4.2
29
Si MAS (7.1 T, v
R
= 7.0 kHz) and
29
Si{
19
F} CP/MAS NMR spectra (7.1 T, v
R
= 10.0 kHz) of a
Portland cement hydrated for 28 days, prepared from a fluoride-mineralized clinker containing 0.32 %w F. The
spectra were recorded with relaxation delays of 30 s for (a) and 4 s for (b). The total numbers of scans for the
experiments are 2048 scans and 62400 scans, respectively. Furthermore, the
29
Si{
19
F} CP/MAS NMR experiment
used a CP contact time of 5.0 ms.
Chapter 4. Fluoride-ion environments in CSH 85

Figure 4.3 (a, c)
29
Si MAS (7.1 T, v
R
= 7.0 kHz) and (b, d)
29
Si{
19
F} CP/MAS NMR spectra (7.1 T, v
R
= 10.0 kHz)
of two synthetic CSH samples prepared with Ca/Si molar ratios of 1.25 and 0.83, respectively. The F/Si molar
ratio in both samples is 0.3. The spectra were recorded with relaxation delays of 30 s for (a, c) and 10 s for (b, d).
The total numbers of scans for the experiments are 2048 scans and 22400 scans, respectively. Furthermore, the
29
Si{
19
F} CP/MAS NMR experiment used a CP contact time of 5.0 ms.

In contrast to the hydrated Portland cement sample, the synthetic CSH samples exhibit
29
Si MAS NMR spectra with three well-resolved peaks at 79.2 ppm, 83.3 ppm and 85.2 ppm,
Figure 4.3. Furthermore, the broad resonance covering a spectral region from 90 ppm to 95
ppm, Figure 4.3 (c), may be assigned to a Q
3
site, which is the characteristic chain-branching
silicate of the tobermorite structure
[165,166,168]
. According to previous studies, the resonances at
85.2 ppm may account for both Q
2
B
and Q
2
P
units while the resonance with o(
29
Si) = 79.2
ppm should be attributed to the chain-end Q
1
or dimeric silicate species. The resonance at about
83 ppm has been identified in several studies
[63,168,169]
and it is assigned to the single protonated
bridging silicate, referred to as Q
2
B*

in Figure 4.1; however, this assignment has not yet been
fully proven experimentally. Additionally, the
29
Si{
19
F} CP/MAS NMR spectra reveal a
resonance of low intensity at 73.9 ppm, which may originate from monomeric silicate species,
Q
0
(H), as a result of the edging effect
[170]
.
To achieve further information about the site preference of the fluoride guest ions and
their influence on the formation of the CSH structures, a series of synthetic CSH samples
including molar ratios Ca/(Si + Al) = 0.83 1.75, where Al/Si = 0 and 0.05, and F/Si = 0 0.5
have been studied by
19
F and
29
Si MAS NMR, with the main findings presented below.
Chapter 4. Fluoride-ion environments in CSH 86
4.2.1. Site preferences of F

ions in the CSH structure from


19
F MAS NMR
19
F MAS NMR spectra of a synthetic CSH sample recorded at two different magnetic
fields, i.e. 7.1 T and 14.1 T, are shown in Figure 4.4. The spectra reveal two distinct fluoride
environments with isotropic chemical shifts at 122.0 ppm and 101.4 ppm. In fact, the same
resonances are observed for all of the synthetic CSH samples, although they exhibit different
spectral intensities as a result of the variation in the Ca/Si molar ratio. The narrow peak at
122.0 ppm may originate from an ordered fluoride structure which exhibits a large CSA pattern
covering a spectral width of nearly 250 ppm. This CSA pattern can be simulated by including
two different
19
F sites with identical isotropic chemical shift of 122.0 ppm where one has a
large chemical shift anisotropy (o
o
) and the other possesses a small CSA. However, the
optimized simulations for the spectra recorded at 7.1 T and 14.1 T result in two distinct sets of
CSA parameters, Figure 4.4. This observation indicates that these fluoride ions are located in
nearly identical local structures with an ordered arrangement but with a slight distribution in
their CSA tensors. The second
19
F resonance observed at 101.4 ppm is shifted slightly towards
higher frequency relative to the resonance observed for CaF
2
, o(
19
F) = 105.9 ppm. Furthermore,
it exhibits a symmetric spinning sideband pattern covering a spectral region from 180 ppm to
20 ppm (Figure 4.5), similar to that for CaF
2
, indicating that this resonance is affected by strong
dipolar couplings such as
19
F
19
F or
1
H
19
F. Furthermore, the rather large line width of the
centerband for this resonance indicates that these fluoride ions occur in a disordered local
environment.
According to the structure model for CSH proposed by Richardson and Groves
[62,170]
,
two distinct types of hydroxyl groups are available for replacement with F

ions, i.e. the


covalently bonded hydroxyl groups of the Q
1
and Q
2
B
sites and the free OH

ions distributed in
the interlayer of the T-based part or in the principal layer of the J-based part of CSH structure;
those are sketched in Figure 4.1. In order to obtain information about the SiF connectivities in
the CSH structure, the
29
Si{
19
F} CP hetero-nuclear correlation (HETCOR) experiment was
applied to a synthetic CSH sample containing Ca/Si = 1.25 and F/Si = 0.3. However, a
reconstruction of the 2D spectrum has not been possible due to a short T
2
relaxation for the
19
F
spins, which causes the
19
F magnetization to dephase completely within very short time under
the t
1
evolution. Alternatively, the
29
Si{
19
F} FBCP experiment, described in section (2.2.7.), was
applied to identify
19
FSi connectivities. Since this experiment utilizes the dipolar coupling to
transfer magnetization between the
19
F and
29
Si spins, the absence of a resonance at 122.0 ppm
in the
29
Si{
19
F} FBCP spectrum, Figure 4.5 (b), indicates that these fluoride ions are located far
Chapter 4. Fluoride-ion environments in CSH 87
away from the dreierketten silicate chains. Thus, they are tentatively assigned to F

ions that
substitute for the OH

groups in the principal layer of the J-based CSH structure. This


assignment is supported by the large chemical shift anisotropy observed for this resonance,
which may be a result of the ordered local structure in the principal layer.


Figure 4.4
19
F MAS NMR spectra of a synthetic CSH sample prepared with a Ca/Si molar ratio of 1.75 and F/Si
= 0.3. Spectrum (a) was recorded at 7.1 T using a spinning speed v
R
of 10.0 kHz, a 10-s relaxation delay and 128
scans. Spectrum (b) was recorded 14.1 T with v
R
= 13.0 kHz, a 15-s relaxation delay and 512 scans. The
corresponding simulated spectra using the STARS program are shown in (b) and (d), respectively. The simulations
included three
19
F sites one with isotropic chemical shift at 101.4 ppm (o
o
= 0) and two with identical isotropic
chemical shifts of 122.0 ppm. However, the refined simulations (root-mean-square, rms < 1 %) of the spectra
result in two different set of CSA parameters; these are o
o
= 132 ppm and 48 ppm, q
o
= 0.02 and 0.08 with relative
intensity ratio of 1.0 : 3.1 for (b) and o
o
= 112 ppm and 57 ppm, q
o
= 0.13 and 0.34 with a relative intensity ratio
of 1.0 : 2.4 for (d).


Chapter 4. Fluoride-ion environments in CSH 88

Figure 4.5
19
F MAS (a) and
19
F{
29
Si} FBCP MAS (b) NMR spectra (7.05 T, v
R
= 10.0 kHz) of a synthetic CSH
sample prepared using a Ca/Si molar ratio of 1.25 and F/Si = 0.3. Both spectra were recorded using a 10-s relaxation
delay. The spectrum in (a) is achieved with 512 scans while (b) is recorded with 58482 scans. Furthermore, the
19
F
29
Si cross polarization transfers were achieved using an 8-ms CP contact time.

The resonance at o(
19
F) = 101.4 ppm, which exhibit a
19
F
29
Si connectivity (Figure 4.5),
may originate either from F atoms that substitute for the covalently bonded OH groups of the Q
1

and Q
2
B
silicates or from the fluoride ions distributed in the interlayer of the T-based CSH
structure. In this particular case, the
29
Si{
19
F} CP-REDOR experiment (Section 2.2.8) may be
applied to determine the average SiF distances for each of the Q units, facilitating an
unambiguous assignment. However, due to the limited time available for this project, such
experiments have not been conducted. Nevertheless, recent theoretical calculations as well as
experimental investigations of the incorporation of fluoride in bioactive calcium/alkali-silicate
glasses demonstrate that SiF covalent bonds in tetrahedral environments is unlikely to form in
such systems, which is consistent with the absence of resonance from the Q
2
B*

component
(o(
29
Si) = 83.3 ppm) in the
29
Si{
19
F} CP/MAS NMR spectra shown in Figure 4.3. Therefore,
the
19
F resonance at 101.4 ppm is most likely associated with the fluoride ions distributed in the
interlayer of the T-based CSH structure. This assignment is supported by the observation of a
large linewidth and a large symmetric spinning sideband pattern for the
19
F resonance, which
may result from the disordered structure of the interlayer and the strong dipolar couplings
between these fluoride ions and H atoms from the OH

groups or the water molecules.


Chapter 4. Fluoride-ion environments in CSH 89
4.2.2. Influence of fluoride guest ions on the CSH structure
29
Si MAS NMR spectra for a series of synthetic CSH samples, illustrating the effect of
fluoride on the polymerization of the silicate chains, are shown in Figure 4.6. The CSH
samples were prepared from raw mixes including a Ca/Si molar ratio of 1.25 and F/Si molar
ratios in the range 0 0.5. The
19
F MAS NMR experiments for these samples only show the two
resonances at 101.4 ppm and 122.0 ppm which indicates that at the studied concentrations,
fluoride are exclusively incorporated in the CSH phase. However, the distribution of fluoride
ions between these two fluoride environments exhibits variations with increasing F/Si molar
ratio, Figure 4.7. Overall, the increased bulk fluoride content seems to increase the amount of
fluoride guest ions in the T-based CSH structure, reflected by the decreased I
[-122 ppm]
/I
[101 ppm]

ratio. However, a slight increase is observed at F/Si = 0.5.


Figure 4.6
29
Si MAS NMR spectra (7.1 T, v
R
= 7.0 kHz) of synthetic CSH samples synthesized with a Ca/Si
molar ratio of 1.25 and F/Si molar ratios of (a) 0, (b) 0.1, (c) 0.2, (d) 0.3 and (e) 0.5. The spectra were recorded
using a 30-s relaxation delay and typically 2048 scans.
Chapter 4. Fluoride-ion environments in CSH 90
As it can be seen from the overall features of the
29
Si MAS NMR spectra shown in Figure 4.6,
the increased bulk fluoride content leads to a slight decrease in the line width for all
29
Si
resonances indicating that a CSH with a higher locally ordered structure is formed. Moreover,
it significantly increases the amount of the Q
2
components at the expense of Q
1
units. The
average silicate chain length, <CL> = (Q
2
+ Q
1
)/Q
1
, as a function of the F/Si molar ratios is
determined from deconvolutions of the
29
Si MAS NMR spectra and plotted in Figure 4.7. For all
spectra, in order to obtain acceptable lineshape simulations, it includes two Q
1
resonances (79
ppm and 80 ppm) and two Q
2
resonances (83 ppm and 85 ppm) whereas the Q
0
(H) resonance
at 74 ppm has been neglected. The plot of <CL> versus the F/Si molar ratio clearly shows that
for a fixed Ca/Si molar ratio an increased incorporation of fluoride ions in the CSH phase
tends to promote the polymerization of the SiO
4
tetrahedra. In addition, the chemical shifts of the
Q
1
and Q
2
sites experience a small shift towards more negative ppm values with increasing F/Si
molar ratio, Figure 4.8.


Figure 4.7 (Left) Plot of the I
[122 ppm]
/I
[101 ppm]
ratio, where I is the spectral

intensities of the
19
F resonances at 122
ppm and 101 ppm, as a function of the F/Si molar ratio. The plot reflects the distribution of fluoride ions in the T-
based part relative to J-based part of CSH structure. (Right) Plot of the average silicate chain length (i.e. <CL> =
(Q
1
+ Q
2
)/Q
1
) for the CSH versus the F/Si molar ratio. The data are obtained from
19
F MAS and
29
Si MAS
NMR experiments at 7.1 T, respectively.

Chapter 4. Fluoride-ion environments in CSH 91

Figure 4.8 Plots of
29
Si isotropic chemical shifts (o
iso
) for the Q
1
and Q
2
sites as a function of the F/Si molar ratio
(left) and the Ca/(Si + Al) molar ratio (right). The values of o(
29
Si) are determined from
29
Si MAS NMR
experiments at 7.1 T. Furthermore, the graph shown on the right includes
29
Si chemical shifts for two series of
synthetic CSH samples, i.e. with (Al/Si = 0.05) and without including Al
2
O
3
which are shown in red and blue,
respectively.

Similar trends are observed for two other series of synthetic CSH samples, which are
synthesized with and without including an aluminum source (Al/Si = 0.05). Furthermore, the
syntheses used a fixed F/Si molar ratio of 0.3 but with Ca/Si ratios varying from 1.75 to 0.83; the
Ca/Si ratio of 1.75 is the typical average value for CSH in mature cement pastes while 0.83
and 1.5 are the ideal values for tobermorite 14- and jennite, respectively. The results from the
19
F MAS and
29
Si MAS NMR experiments for these samples are summarized in Figure 4.8 and
Figure 4.9. As it can be seen from Figure 4.8, the increase in the Ca/Si molar ratio slightly shifts
the Q
1
and Q
2
resonances towards positive ppm values. In fact, the same trend is reported in
previous investigations
[169,171,172]
, although for CSH without including fluoride. Evidence for
the incorporation of Al
3+
in the tetrahedrally coordinated environment is obtained from
27
Al
MAS NMR experiment on a synthetic CSH sample (Ca/Si = 1.25), Figure 4.10. The spectrum
clearly shows two resonances in the spectral region for Al
3+
ions in tetrahedral coordination, i.e.
67 ppm and 73 ppm. However, their total spectral intensity only account for half of the bulk
Al
2
O
3
content. The remainding Al
2
O
3
occurs as octahedrally coordinated Al species such as the
third aluminate phase
[173]
(4 ppm) and AFm phases (9 ppm). A minor amount of aluminum also
occurs in penta-coordinated environments (36 ppm). As the
29
Si chemical shift primarily reflects
local structural features within the first and second coordination sphere, this Al
3+
incorporation
Chapter 4. Fluoride-ion environments in CSH 92
in the bridging sites only introduces a shift of approximately 3 ppm towards positive ppm value
on the two Q
2
P
silicates, which are referred to as Q
2
(1Al) in Figure 4.1, whereas the remaining Q
1

and Q
2
resonances (including Q
2
B
and Q
2
P
) are nearly unaffected by presence of the Al
3+
guest
ions, Figure 4.8.
A comparison of the reciprocal average silicate chain length for the two series of samples
with a previously reported data set
[174]
for a series of CSH sample without fluoride ions is
shown in Figure 4.9. Consistent with earlier observations, this figure shows that a decrease in the
Ca/(Si + Al) molar ratio results in an increase in the average silicate chain length. However, in
the presence of fluoride guest ions, CSH phases with much longer average silicate chains are
formed. The effect of fluoride guest ions on the average chain length seems to be increased with
decreasing Ca/(Si + Al) molar ratio. For Ca/(Si + Al) ratios above 1.25, the average silicate chain
length of the fluoride-containing CSH phases tends to be increased by the presence of an
aluminum source whereas it is reduced below this value.
The distribution of fluoride ions between the J-based and T-based CSH structures is
also affected by the presence of Al
3+
guest ions, although a significant difference is only
observed for samples with Ca/(Si + Al) ratios of 1.25 and 1.5. For all samples, only two
19
F
resonances with isotropic chemical shifts at 101.4 ppm and 122.0 ppm are observed.
However, the distribution of fluoride over these two sites has shown a strong dependency on the
molar ratios Ca/(Si + Al), Al/Si as well as F/Si. It is apparent from a plot of the I
[-122 pm]
/I
[-101 ppm]

ratio as a function of the Ca/(Si + Al) molar ratio that the ordered fluoride environment is
predominant in samples with high bulk CaO content. As the Ca/(Si + Al) ratio is decreased, the
resonance at 101.4 ppm exhibits an increasing intensity. The most significant change is
observed when the Ca/(Si + Al) ratio is reduced just below 1.5, i.e. the ideal Ca/Si molar ratio of
jennite. This observation strongly supports the assignment of the
19
F resonances, o(
19
F) = 101.4
ppm and 122.0 ppm, to fluoride ions incorporated in the T-based and J-based CSH
structures, respectively. In the absence of an aluminum source, the intensity of the resonance at
101.4 ppm becomes much larger than that at 122.0 ppm as the Ca/Si molar ratio is decreased
to 1.25. However, below this value, the I
[101 ppm]
/I
[122 ppm]
ratio is slightly increased and the
distribution of fluoride in the two different environments become almost equal as the Ca/Si
molar ratio is further reduced to 0.83. This observation is consistent with the fact that the J-based
CSH structure is predominant at high Ca/Si ratios and therefore, a larger amount of fluoride
ions are incorporated into this phase. As the Ca/Si ratio decreases, a larger amount of T-based
structure will be formed within the CSH phase leading to an increased intensity for the
fluoride resonance at 101.4 ppm.
Chapter 4. Fluoride-ion environments in CSH 93

Figure 4.9 (Left) Plot of the reciprocal average silicate chain length, <CL>
1
, as a function of the Ca/(Si + Al) molar
ratio for three series of synthetic CSH samples: () F/Si = 0.3 and Al/Si = 0, () F/Si = 0.3 and Al/Si = 0.05, ()
F/Si = 0 and Al/Si = 0.05. (Right) Plot of the I
[-122 ppm]
/I
[-101 ppm]
ratio, where I is the spectral

intensitiy of the
19
F
resonances at 122.0 ppm and 101.4 ppm, as a function of the Ca(Si + Al) molar ratio.

A possible explanation for the similarity in the shifting effects on the Q
1
and Q
2

resonances as a result of an increase in the fluoride content and the decreased Ca/Si molar ratios
(Figure 4.8) is that the increased incorporation of fluoride in the T-based CSH structure is
accompanied by a similar increase in the quantity of Ca
2+
ions in the interlayer of this phase.
This effectively reduces the Ca/Si molar ratio for the main calcium layer and therefore, it
increases the average silicate chain and promotes the formation of the T-based CSH structure
at the same time. Furthermore, the absence of the Q
2
B*
resonance in the
29
Si{
19
F} CP/MAS
spectra shown in Figures 4.2 and 4.3 indicates that these fluoride guest ions most likely
substitute for the OH

groups associated with CaOH on the bridging sites of the silicate chains
rather than distributed uniformly in the interlayer. Thus, a large amount of fluoride ions may be
incorporated in the dimeric T-based CSH structure, which may predominate for Ca/Si ratios
just below the ideal value of the jennite structure (i.e. 1.5). As the Ca/Si molar ratio is further
decreased, SiO
4
dimers polymerizes to form longer silicate chains (Figure 4.9), reducing the
amount of Ca
2+
ions on the bridging sites and the quantity of fluoride guest ions in the T-based
CSH structure at the same time. A similar effect may result from the incorporation of Al
3+
in
the bridging sites. Furthermore, as it is apparent from Figure 4.10, the formation of calcium
aluminate hydrate phases may reduce the Ca/(Si + Al) molar ratio for the CSH phase and
Chapter 4. Fluoride-ion environments in CSH 94
therefore, it affects the average silicate chain length as well as the distribution of fluoride
between the T-based and J-based CSH structures.


Figure 4.10
27
Al MAS NMR spectrum (14.1 T, v
R
= 13.0 kHz) of a synthetic CSH sample prepared using a
Ca/(Si + Al) molar ratio of 1.25 and Al/Si = 0.05. The fluoride content used in the synthesis corresponds to F/Si =
0.3. The spectrum was recorded using a 0.5-s excitation pulse, a 2-s relaxation delay and 27520 scans. The
tetrahedrally coordinated Al
3+
guest ions of the CSH phase are identified by the two resonances at about 73 ppm
and 67 ppm. The Al
3+
ions in penta-coordination appear at about 36 ppm while the octrahedral aluminum species,
the third aluminate phase and AFm, are observed as a sharp peak at 4 ppm and a shoulder at 9 ppm, respectively.



Chapter 4. Fluoride-ion environments in CSH 95
4.3. Influence of fluoride on the hydration of Portland cement
Based on the observations from the
29
Si and
27
Al NMR experiments of the hydration of
fluoride-mineralized Portland cements presented in Section (3.5), the hydration of the
unmodified white Portland cement (0.04 %w F) and three selected fluoride-mineralized cements,
containing approximately 0.04, 0.29 and 0.9 %w F, have been examined using
19
F,
27
Al and
29
Si
MAS NMR. Hydrated samples with ages from 6 hours up to 200 days have been prepared using
the procedure described in Appendix 1.

4.3.1. Identification of CaF
2
in hydrated Portland cement by
19
F MAS NMR
The ability to identify CaF
2
formation during the early hydration of Portland cement may
be decisive for the understanding of the retarding mechanism of fluoride. The
43
Ca{
19
F}
CP/MAS NMR experiment, described in Section (2.2.4.), may provide a clear distinction for the
resonance from Ca
2+
ions of the insoluble CaF
2
salt from the remaining calcium-containing
phases. However, due to the very low CaF
2
content in combination with the low natural
abundance of the
43
Ca nuclear spin isotope, it has so far not been possible to detect the
43
Ca
signal from this phase for hydrated Portland cement samples.
Figure 4.11 demonstrates an alternative approach for the identification of CaF
2
in
hydrated Portland cement using
19
F MAS NMR. The experiment utilizes the fact the
19
F nuclear
spins of CaF
2
has a much longer spin-lattice relaxation time than that for F

ions present in the


CSH phase; the required repetition time for complete spin-lattice relaxation of the
19
F spins in
pure CaF
2
is about 120 s at 14.1 T while it is only 15 s for the fluoride guest ions in CSH.
Thus, the
19
F MAS NMR spectrum recorded using a 120-s relaxation delay may contain
19
F
resonances from CaF
2
(o(
19
F) = 106 ppm) as well as the CSH phases (o(
19
F) = 103 ppm
and 122 ppm). In a second experiment where the
19
F MAS NMR spectrum is recorded using a
relaxation delay of only 15 s, the
19
F resonance of CaF
2
only contributes to the first few scans
and will be partly saturated afterwards. As demonstrated in Figure 4.11 for a Portland cement
(0.9 %w) which has been hydrated for 200 days, the difference spectrum corresponds to the
spectrum of CaF
2
. For this particular sample, the subtraction of spectra recorded with different
relaxation delay seems not to be necessary. However, as demonstrated in the following section,
the CaF
2
content in samples corresponding to one-day of hydration is much lower and therefore,
the subtracting procedure is necessary in order to identify the
19
F resonance from CaF
2
.

Chapter 4. Fluoride-ion environments in CSH 96

Figure 4.11
19
F MAS NMR spectra (14.1 T, v
R
= 13.0 kHz and 2048 scans) of a Portland cement (0.9 %w F)
hydrated for 200 days. The spectrum shown in (a) was recorded using a 120-s relaxation delay whereas it was 15 s
in (b). The difference spectrum (c) shows the
19
F resonance from the insoluble CaF
2
salt. The asterisks (*) indicates
the spinning sidebands from the CaF
2
phase.


4.3.2. Hydration of fluoride-mineralized Portland cement studied by
19
F MAS NMR
As it is demonstrated in Figure 4.12 4.15, the high sensitivity of
19
F MAS NMR makes
it possible to follow the hydration of fluoride ions in Portland cements, even for very small
fluoride concentration of only 0.04 %w F, within reasonable spectrometer times; the spectra
presented in Figure 4.12 and 4.13 was obtained using a 8-s relaxation delay and typically 8190
scans which corresponds to approximately 18 hours whereas the spectra in Figure 4.14 and 4.15
can be obtained within two hours applying a repetition time of 8 s and typically 512 scans. Since
fluoride is entirely incorporated in the alite phase, the hydration of fluoride may somehow reflect
the hydration of this phase. In general, the
19
F MAS spectra of hydrated Portland cements
contain resonances from the anhydrous phase centered at 114 ppm and two resonances at 122
ppm and 103 ppm, which have been identified as fluoride guest ions of the CSH phases as
described above. These resonances experience variation in their intensities as the hydration
proceeds. Furthermore, it can be seen that the release of fluoride ions from the anhydrous phase
is nearly complete after 28 days of hydration. In addition, the insoluble CaF
2
salt (o(
19
F) = 106
Chapter 4. Fluoride-ion environments in CSH 97
ppm) has been identified in samples prepared from the modified cements containing 0.29 and 0.9
%w F after hydration for one day, Figure 4.16.
Considering the two series of samples shown in Figure 4.12 and 4.13, they contain
approximately the same fluoride concentration, but differ significantly in their Ca/Si ratios and
Al
2
O
3
contents; the unmodified white Portland clinker has a fluoride content of 0.04 %w, 70.30
%w CaO, 25.42 %w SiO
2
and 2.13 %w Al
2
O
3
while the modified clinker has 0.04 %w, 68.40
%w CaO, 24.98 %w SiO
2
and 4.43 %w Al
2
O
3
. Thus, apart from the difference in the distribution
of fluoride guest ions between the T- and J-based CSH structures, which is possibly due to the
difference in their Ca/Si ratios, the fluoride species in the two cements show very similar
hydrational properties. Both cements show clear hydrational activities at 12 hours after mixing
with water, indicated by the decrease in intensity of the resonance at 114 ppm and the
appearance of the resonance at 122 ppm. At hydration ages up to seven days, the resonance at
122 ppm exhibits the predominant intensity while the fluoride guest ions of the T-based CSH
structure is only present in small amount. As the hydration proceeds, the resonance at about
102 ppm exhibits an increasing intensity with time whereas the fluoride guest ions of the J-
based CSH structure experience variation in quantity and finally decrease to nearly zero after
200 days of hydration. This is consistent with earlier observations for the CSH structures
formed from the hydration of Portland cement
[58,175]
, which identify tobermorite- as well as
jennite-like structures for the CSH formed during the early hydration period while after longer
hydration times, the CSH phase turns out to be suitably described by the T/CH viewpoint. For
both cements, the fluoride guest ions in alite are nearly consummed after 28 days of hydration.

Chapter 4. Fluoride-ion environments in CSH 98

Figure 4.12
19
F MAS NMR spectra (7.1 T, v
R
= 10.0 kHz) of hydrated samples prepared from the original white
Portland clinker (0.04 %w F) with the hydration times given below each spectrum. The spectra were recorded using
an 8-s relaxation delay and typically 8192 scans. The
19
F resonance from the anhydrous phase is observed at 114
ppm whereas the fluoride guest ions from the CSH phases appear at 102 ppm and 122 ppm.


Figure 4.13
19
F MAS NMR spectra (7.1 T, v
R
= 10.0 kHz) of hydrated samples prepared from a modified white
Portland clinker (0.04 %w F and 4.3 %w Al
2
O
3
) with the hydration times given below each individual spectrum.
The spectra were recorded using a 8-s relaxation delay and typically 8192 scans.
Chapter 4. Fluoride-ion environments in CSH 99

Figure 4.14
19
F MAS NMR spectra (7.1 T, v
R
= 10.0 kHz) of hydrated samples prepared from a modified white
Portland clinker (0.29 %w F and 4.3 %w Al
2
O
3
) with the hydration times given below each individual spectrum.
The spectra were recorded using a 8-s relaxation delay and typically 512 scans.



Figure 4.15
19
F MAS NMR spectra (7.1 T, v
R
= 10.0 kHz) of hydrated samples prepared from a modified white
Portland clinker (0.9 %w F and 4.3 %w Al
2
O
3
) with the hydration times given below each individual spectrum. The
spectra were recorded using a 8-s relaxation delay and typically 512 scans.
Chapter 4. Fluoride-ion environments in CSH 100
Fluoride in the cement prepared from the clinker containing 0.29 %w F exhibits a
somewhat slower hydration rate within the first 12 hours after mixing with water; only minor
changes are observed in the
19
F MAS NMR spectra recorded for the hydrated sample stopped
after 12 hours as compared to that of anhydrous cement. For longer hydration times, it shows
similar hydration features to the two cements described above. After 24 hours of hydration, the
samples clearly show a large proportion of fluoride ions incorporated in both the T- and J-based
CSH structures, i.e. the clear appearances of resonances at 102 ppm and 122 ppm,
respectively. Moreover, a minor amount of CaF
2
salt is formed at this hydration stage, identified
by the appearance of the
19
F resonance at 106 ppm, Figure 4.14. It is clearly evidenced from the
experiments that the quantity of CaF
2
observed after one day of hydration increases with
increasing fluoride content in the anhydrous cement.
As the fluoride content in the cement is increased much above the critical value (i.e. 0.36
%w F in cement clinker), the alite phase seems not to react with water before 12 hours; the
19
F
MAS NMR spectrum of this sample closely resembles that of the anhydrous cement, Figure
4.15. Moreover, instead of being incorporated in the CSH phase, a large amount of fluoride
ions precipitates as CaF
2
within the first 12 24 hours after mixing, Figure 4.16. Furthermore,
the absence of
19
F resonances at 122 and 102 ppm reflects that only a minor amount of
fluoride ions is incorporated in the CSH phase at this hydration stage.


Figure 4.16
19
F MAS NMR spectra (7.1 T, v
R
= 10.0 kHz) of a synthetic CSH sample (Ca/Si = 1.25 and F/Si =
0.5) and samples hydrated for one day, prepared from the three modified cements with fluoride contents shown
below their corresponding spectra. The difference spectra clearly demonstrate that CaF
2
is only formed in the
hydrated modified cements, where the corresponding clinkers contain 0.29 %w F and 0.90 %w F, respectively.

Chapter 4. Fluoride-ion environments in CSH 101
4.3.3. Retarding mechanism of fluoride ions on the hydration of Portland cement
An understanding of the retarding mechanism of fluoride on the hydration of Portland
cement may be derived when evaluating the results from the
19
F,
27
Al and
29
Si MAS NMR
experiments together with observations from previous investigations
[155,162,164]
. It is apparent
from the
27
Al MAS NMR spectra shown in Figure 4.18 that a significant amount of AFt phases
(e.g. ettringite) is already formed within the first six hours after mixing, regardless the fluoride
content. From six to twelve hours after mixing, the hydration of the aluminum species in
cements with fluoride contents below the critical value (i.e. 0.36 %w F) proceeds, although with
a very slow hydration rate. The conversion of AFt into AFm phases is already observed for the
modified cement containing 0.04 %w F after 12 hours of hydration while in the modified cement
containing 0.29 %w F, the formation of ettringite is still proceeding. On the other hand, the
fraction of aluminum in the anhydrous and hydrated phases of the modified cement containing
0.9 %w F exhibits no significant changes from six to twelve hours of hydration. In conjunction
with the observations by
19
F MAS NMR, summarized in Figure 4.15, this strongly indicate that
the alite phase in the cement containing 0.9 %w F is intact within this hydration period. This
retarding effect might be an intrinsic property of the fluoride-mineralized alite
[155]
as proposed
earlier since fluoride is not yet released from the alite phase and thus, CaF
2
cannot be formed at
this hydration stage.


Figure 4.17 The degree of hydration for alite and belite in fluoride-mineralized cements containing different
quantities of fluoride: () 0.04 %w F, () 0.29 %w F and () 0.90 %w F. The relative fractions of the silicate
species are obtained from deconvolutions of the
29
Si MAS NMR spectra (7.1 T, v
R
= 7.0 kHz, 30-s relaxation delay)
of the hydrated samples, stopped at different hydration times up to 100 days.
Chapter 4. Fluoride-ion environments in CSH 102


Figure 4.18
27
Al MAS NMR spectra (14.1 T, v
R
= 13.0 kHz) of selected hydrated samples for three modified
cements with the fluoride contents shown below the spectra. The tetrahedrally coordinated Al
3+
guest ions of the
alite and belite phases are identified by the resonance at about 82 ppm with a shoulder at 84 ppm, respectively. The
tricalcium aluminate phase appears as a broad resonance from 30 to 85 ppm. The hydrated phases including the AFt
and AFm phases are observed at 13 ppm and 9 ppm, respectively. The asterisks (*) indicate spinning sideband from
the AFt phase.

For one day of hydration and later, the presence of fluoride in quantities above the critical
value tends to affect the hydration of the entire cement. As demonstrated in Section 3.5, the
quantity of fluoride ions strongly modifies the release of Al
3+
from the alite phase and the
conversion of AFt into AFm phases. The examination of the effect of fluoride on the hydration
properties of the alite and belite phases of Portland cement using
29
Si MAS NMR also
demonstrates that below the critical value, the presence of fluoride ions only slightly affects the
hydration of alite and belite, Figure 4.17; up to seven days, the presence of fluoride seems to
increase the reactivity of alite, but it slightly decreases the degree of hydration of belite. When
present in concentrations above the critical value, fluoride has shown a substantial retarding
effect on the hydration of alite as well as belite. It significantly reduces the degree of hydration
Chapter 4. Fluoride-ion environments in CSH 103
of alite up to 28 days. However, for longer hydration times, the quantity of fluoride has only a
marginal effect on the hydration of alite. According to the fact that fluoride is only incorporated
in the alite phase but seems to affect the hydration of the entire cement when present in
quantities above the critical content, a plausible explanation for this is the formation a protective
layer of insoluble CaF
2
salt on the cement grain surface, which prevents the cement to
hydrate
[162,164]
. The formation of such layer is supported by the observation of CaF
2
in the
hydrated fluoride-mineralized cement, Figure 4.16.


4.4. Summary
Site preferences of fluoride ions and their influence on the structure of CSH has been
investigated for synthetic CSH and hydrated fluoride-mineralized Portland cements using
19
F,
27
Al and
29
Si MAS NMR. These studies demonstrate that
19
F MAS NMR represents a unique
tool for the structural characterization of fluoride ions in Portland cement. The high sensitivity of
the
19
F spins makes it possible to follow the hydration of fluoride species even when they are
present in quantities of as low as 0.04 %w F.
From the study of synthetic CSH, it is demonstrated that for the studied fluoride
contents (F/Si = 0 0.5), fluoride ions are almost exclusively incorporated in the CSH
structure. The insoluble CaF
2
salt is not formed in such systems. The
19
F MAS NMR
experiments reveal two different structural environments for the fluoride guest ions of the
CSH phases. The first exhibits a
19
F resonance at o(
19
F) = 122.0 ppm with a large CSA
spinning sideband pattern, which reflects an ordered local structure and therefore, it has been
assigned to fluoride ions distributed in the principal layer of the jennite-based CSH structure.
The second type of fluoride ions appear at o(
19
F) = 101.4 ppm. It is evident from the
19
F{
29
Si}
FBCP/MAS experiment that this resonance is dipolar-coupled to the
29
Si spins from the
dreierketten silicate chains and therefore, it most likely originates from the fluoride ions
incorporated in the interlayer for the tobermorite-based CSH structure. This is supported by
the observation of a symmetric spinning side band pattern for this resonance, which covers
nearly 160 ppm, indicating that these fluoride ions possess strong
1
H
19
F dipolar couplings,
possibly to the H atoms from the OH groups or the water molecules.
The incorporation of fluoride ions in the CSH phase results in an increase in its
average silicate chain length. This effect of fluoride tends to increase with decreasing Ca/Si
ratios. Furthermore, it also seems to be affected by the simultaneous incorporation of Al
3+
ions in
the bridging site of the silicate chain. For Ca/Si ratios above the ideal Ca/Si ratio of the jennite
Chapter 4. Fluoride-ion environments in CSH 104
structure (i.e. Ca/Si = 1.5), the presence of Al
3+
guest ions appears to increase the effect from
fluoride ions on the average silicate chain length whereas below this value, it reduces the
average silicate chain length of the fluoride-containing CSH structure. It is also observed that
the increased incorporation of fluoride ions introduces a slight high-field shift, i.e. towards more
negative o(
29
Si) value, of the
29
Si isotropic chemical shifts for the Q
1
and Q
2
silicates.
The hydration of fluoride ions in Portland cement has been followed by
19
F MAS NMR
for four cements: a commercial white Portland cement and three modified cements prepared
from clinker with a high bulk aluminium content (~ 4.3 %w Al
2
O
3
) and fluoride contents of
0.04, 0.29 and 0.90 %w F. The experiments show that the majority of fluoride is incorporated in
the CSH phases, although the fraction of fluoride distributed between the jennite-based and
the tobermorite-based CSH structures experience variation with hydration time. For hydration
ages up to seven days, the fluoride ions of the jennite-based CSH is predominant while on
longer hydration times, the fluoride ions are preferentially incorporated in the tobermorite-based
CSH structure. For the cements prepared from the modified clinker with fluoride contents of
0.29 and 0.9 %w F, the CaF
2
salt is formed already after one day of hydration. The content of
this phase tends to increase as the hydration proceeds. However, it only constitutes a minor part
of the bulk fluoride content.
The hydration of the cements has also been followed by
27
Al and
29
Si MAS NMR. The
experiments for samples with hydration ages up to 12 hours reveal that the prolonged setting
time might be an intrinsic property of the fluoride-mineralized alite, since the fluoride is not yet
released at this hydration stage and therefore, the CaF
2
cannot be formed. On the other hand,
after one day of hydration, a significant amount of CaF
2
is detected by
19
F MAS NMR for the
cement containing 0.9 %w F. Moreover, the
27
Al and
29
Si MAS NMR experiments demonstrate
that the hydration of all the present phases are affected by the presence of fluoride, despite the
fact that fluoride ions are only incorporated in the alite phase. Therefore, the consideration of the
formation of a protective layer of CaF
2
is plausible for Portland cements with a fluoride content
above the critical value of 0.36 %w F (cf. Chapter 3). This layer prevents the cement grains to
react with water and therefore, it reduces the degree of hydration for all cement phases during
the early hydration period.

Chapter 5. Framework structures of aluminosilicate binders 105










5. Chapter



Framework structures of alumino-
silicate binders from solid-state
NMR spectroscopy










This chapter includes two applications of the
29
Si{
27
Al} REAPDOR NMR experiment in
structural investigations of the network structure of aluminosilicate binders. The first part
concerns a study of the SiOAl connectivities in geopolymeric materials formed from alkali
activation of metakaolin samples. The second part presents the results from a reexamination of
the disordered layer structure of strtlingite using solid-state NMR. More details on both studies
are presented in manuscripts 1 and 2, respectively.
Chapter 5. Framework structures of aluminosilicate binders 106
5.1. SiOAl connectivities in alkali-activated materials
Recent developments of alternative building materials have resulted in a renewed interest
in alkali-activated materials
[176-179]
, often denoted geopolymers, as alternative binders, with
emphasis to the generally much lower energy consumption and CO
2
emission required for
production of these materials as compared to conventional Portland cement
[180]
. Additionally, the
materials provide other advantageous properties like high early strength development, long-term
durability, fire-resistance and storage of hazardous inorganic waste
[181-184]
. The process of
forming the aluminosilicate networks requires activation by a relatively high concentration of
alkali hydroxides, typically NaOH and/or KOH. It involves three main steps
[177,178,185]

(1) Release of monomeric Si(OH)
4
and Al(OH)
4
-
units from the starting materials activated
by alkali hydroxide.
(2) Reorganization and diffusion of ions to form small coagulated structures.
(3) Termination by poly-condensation, i.e., the formation of SiOAl linkages to form a 3-
dimensional framework structure.
The structure of geopolymeric materials as well as their performance is dependent on several
factors, including the concentration of alkali hydroxides, the bulk chemical composition and the
quantity of soluble silicates and aluminates dissolved from the precursors
[186-188]
. Alternatively,
the alkali-activation process can be applied to blended cements
[189,190]
, in which the clinker
substitution level can be increased above the current limit of approximately 30 % by weight.
However, geopolymer cements have only been commercialized in small-scale facilities so far,
but not in large-scale applications where the strength is critical.
In the generally accepted structural model
[185]
, geopolymers are described by a backbone
structure of poly(sialates) with the empirical formula of M
n
{(SiO
2
)
z
AlO
2
}
n
wH
2
O, where the
SiO
4
and AlO
4
tetrahedra are linked by sharing all their oxygen atoms, corresponding to Q
4

units. The presence of cations M such as Na
+
, K
+
, Li
+
, Ca
2+
, Ba
2+
, NH
4
+
and H
3
O
+
located in the
framework cavities is necessary to preserve charge balance of the incorporated aluminium ions
in four-fold coordination. So far, the detailed characterization of those materials has been a
major challenge, owing to the lack of long-range periodicity in their open-framework structures.
Thus, alkali-activated materials are often considered as X-ray amorphous and their XRD
patterns provide very little structural information
[177,185]
.
Chapter 5. Framework structures of aluminosilicate binders 107

Figure 5.1 Experimental (a) and simulated (b)
29
Si MAS NMR spectra of an alkali-activated metakaolin sample
(molar Si/Al = 2.0, Na/Al = 1.5). The experimental spectrum (a) was recorded at 7.1 T using a spinning speed of v
R

= 7.0 kHz and a 30-s relaxation delay. Spectrum (b) is a simulation of (a), in which the eight different resonances
shown in (c) and (d) are included.

According to the general ability of NMR spectroscopy to probe local structural features
independent of long-rang order,
29
Si MAS NMR spectroscopy appears to be a useful tool for
investigations of the open framework structures of alkali-activated materials. However, a
limitation of the single-pulse
29
Si MAS NMR experiment is apparent when a number Al
3+
ions is
substituted into the second-coordination sphere of the probed SiO
4
site; each Al
3+
for Si
4+

Chapter 5. Framework structures of aluminosilicate binders 108
substitution in the second-coordination sphere leads to a resonance shift of about 5 ppm towards
higher frequency
[79]
, causing resonances from different types of SiO
4
condensation to overlap
and preventing a straight-forward interpretation of the observed chemical shifts
[191-194]
. For
example, the chemical shifts of Q
3
(0Al) falls in the same spectral region as those from Q
4
(2Al)
and Q
4
(3Al) sites. This is illustrated for an alkali-activated metakaolin sample in Figure 5.1; the
sample was synthesized using a Si/Al molar ratio of 2.0 and Na/Al molar ratio of 1.5. It is
apparent from the deconvolution, shown in Figure 5.1 (b-d), that the
29
Si MAS NMR spectrum
consists of eight resonances having chemical shifts within the region from 115 to 75 ppm. The
assignment of these resonances from single-pulse
29
Si MAS NMR experiment is somewhat
uncertain, since this chemical shift region may include any of the Q
4
(nAl) and Q
3
(nAl)
components. In such cases, the
29
Si{
27
Al} REAPDOR experiment described in section 2.2.10
appears to be an appropriate tool for probing the SiOAl network. An unambiguous assignment
of the
29
Si resonances may be achieved from a combination of the
29
Si chemical shifts and the
corresponding number of Al atoms in their second-coordination spheres, obtained from
29
Si{
27
Al} REAPDOR experiments.

5.1.1. Effects of Si/Al and Na/Al molar ratios

29
Si MAS NMR spectra for a series of alkali-activated metakaoline samples, synthesized
using Si/Al = 1.0 3.0 and Na/Al = 1.0, are shown in Figure 5.2. As it can be seen from the
spectra, the
29
Si resonances appear in the region from about 115 ppm to 75 ppm, which, in
principel, can be assigned to all types of tetrahedral SiO
4
species. Considering these
29
Si MAS
NMR spectra, a very intense peak at about 85.0 ppm has been observed for the sample prepared
using a Si/Al molar ratio of 1.0. However, four additional
29
Si resonances with isotropic
chemical shifts o(
29
Si) at 78.6 ppm, 83.2 ppm, 87.5 ppm and 92.2 ppm can be identified
from a deconvolution of this spectrum. As the Si/Al molar ratio is increased from 1.0 to 2.0, five
well-separated
29
Si resonances are observed at 87.5 ppm, 92.2 ppm, 97.0 ppm, 102.5 ppm
and 107.1 ppm. A further increase in the Si/Al molar ratio results in locally disordered network
structures, which is indicated by the smooth, broadened lineshape of the
29
Si MAS NMR spectra
shown in Figure 5.2 (d) and (e). Finally, when the Si/Al molar ratio becomes larger than 3.0, a
significant amount of non-reacted metakaolin have been detected by the
29
Si as well as the
27
Al
MAS NMR experiments, cf. manuscript 1.
Chapter 5. Framework structures of aluminosilicate binders 109

Figure 5.2
29
Si MAS NMR spectra (7.05 T, v
R
= 7.0 kHz) of alkali-activated metakaoline samples, prepared using a
Na/Al molar ratio of 1.0 and Si/Al = 1.0 (a), 1.5 (b), 2.0 (c), 2.5 (d) and 3.0 (e). The spectra were recorded
employing a 30-s relaxation delay and typically 2048 scans.

As it is stated above, the use of alkali in a correct amount is vital for the synthesis of
aluminosilicate network structure since the substitution of Al
3+
for Si
4+
on the tetrahedral sites
has to be charge-balanced. Thus, deficient or excessive amounts of alkali may affect the
formation of the aluminosilicate structures substantially, leading to several different SiO
4

environments within the network structure. This is illustrated in Figure 5.3 showing selected
29
Si
MAS NMR spectra for another series of alkali-activated samples, for which the Si/Al molar ratio
is 2.0 and the Na/Al molar ratio is varied from 0.8 to 2.15. However, only samples with Na/Al s
1.5 have been studied; the alkali-activated product becomes gel-like and is very difficult to pack
in a NMR rotor as the Na/Al molar ratio is raised above 1.5. As it is apparent from Figure 5.3,
the spectra consist of partly overlapping
29
Si resonances with isotropic chemical shifts in the
Chapter 5. Framework structures of aluminosilicate binders 110
range from 107 to 79 ppm. Furthermore, the
29
Si resonances that appear at 85.0, 83.2 and
78.6 ppm are only observed with significant intensity for samples including an excessive
amount of Na
+
ions (i.e. Na/Al > 1.0).
Altogether, when changing the raw mix composition (e.g. Na/Al and Si/Al molar ratios)
used in the syntheses, eight
29
Si resonances within a spectral region from 75 ppm to 110 ppm,
i.e. the chemical shift region for silicon in tetrahedral coordination, have been observed for the
alkali-activated metakaolin samples. The assignment of these resonances from their chemical
shifts is somewhat uncertain since this chemical-shift region may include any of the Q
1
, Q
2
(nAl),
Q
3
(nAl) and Q
4
(nAl) components.


Figure 5.3
29
Si MAS NMR spectra (7.05 T, v
R
= 7.0 kHz) of alkali-activated metakaolin samples, prepared using a
Si/Al molar ratio of 2.0 and Na/Al = 0.8 (a), 10.0 (b), 1.3 (c) and 1.5 (d). The spectra were recorded employing a 30-
s relaxation delay and typically 2048 scans.
Chapter 5. Framework structures of aluminosilicate binders 111
5.1.2.
29
Si {
27
Al} REAPDOR NMR
In order to assign the eight
29
Si resonances, observed for alkali-activated metakaolin
samples presented above,
29
Si{
27
Al} REAPDOR NMR has been applied to determine the
number of Al atoms substituted in the second-coordination sphere for each of the resonances.
This study involves three selected alkali-activated metakaolin samples corresponding to the
spectra shown in Figures 5.2 (a), 5.3 (c) and 5.3 (d).
The first sample considered in this study contains Si/Al and Na/Al molar ratios of 2.0 and
1.3, respectively. Its
29
Si MAS NMR spectrum, Figure 5.3, clearly shows five distinct resonances
separated by approx. 5 ppm. Additionally, the
29
Si{
1
H} CP/MAS NMR experiment (not shown)
of this sample reveals a broad resonance of low intensity at about 85.0 ppm. However, this
resonance has been neglected in the deconvolution of the
29
Si{
27
Al} REAPDOR spectra owing
to its very low intensity. It appears from the
29
Si{
27
Al} REAPDOR spectra shown in Figure 5.4
that only four of the
29
Si resonances are affected by the re-introduced SiAl dipolar couplings.
They experience different rates of SiAl dipolar dephase, reflecting that the numbers of Al atoms
incorporated in their second-coordination sphere are different. On the other hand, the resonance
at 107.1 ppm shows no significant change in its intensity (S
0
~ S) and therefore, it may be
assigned to a Q
4
(0Al) unit. The individual
29
Si{
27
Al} REAPDOR spectra were deconvolved to
assess the intensities, S
0
and S, for the different sites. From a plot of the REAPDOR fractions
versus the evolution time, Figure 5.5, it is confirmed that only four
29
Si resonances are affected
by the SiAl dipolar couplings. The kn
Al
values associated with their dipolar dephases are
obtained from the curve fits shown in Figure 5.5 (right). However, it has been necessary to
include AS/S
0
> 0.3 in the curve fit for the resonance at 87.5 ppm, since it has only one data
point that fulfils the condition AS/S
0
0.3. The kn
Al
values are summarized in Table 5.1.
Complementary results have been achieved for another alkali-activated metakaolin
sample corresponding to that with the
29
Si MAS NMR spectrum shown in Figure 5.3(d). From
the spectral deconvolution shown in Figure 5.1, it can be seen that the sample includes eight
non-equivalent
29
Si sites: six
29
Si resonances having the same isotropic chemical shifts as
observed for the spectrum in Figure 5.3(c) and two additional resonances at 78.6 and 83.2
ppm with low intensities. The last two resonances have been neglected in the deconvolution of
the
29
Si{
27
Al} REAPDOR spectra for this sample due two their rather low intensities. The kn
Al

values obtained from curve fits of AS/S
0
> 0.3 for the remaining
29
Si resonances are listed in
Table 5.1.
Chapter 5. Framework structures of aluminosilicate binders 112

Figure 5.4
29
Si{
27
Al} REAPDOR (14.1 T, v
R
= 10.0 kHz) spectra of an alkali-activated metakaolin sample with
Si/Al and Na/Al molar ratios of 2.0 and 1.3, respectively. (Left) Full signal, S
0
, which is not affected by the SiAl
dipolar couplings. (Right) The attenuated signals, S, reflecting the SiAl dipolar couplings. The evolution periods
are (a): 4 T
r
, (b): 8 T
r
, (c): 12 T
r
and (d) 16 T
r
, where T
r
= 0.1 ms.


Figure 5.5
29
Si{
27
Al} REAPDOR curves for an alkali-activated metakaolin sample with Si/Al and Na/Al molar
ratios of 2.0 and 1.3, respectively: () 87.5, () 92.4, (-) 97.2 and ( ) 102.5 ppm. (Left) Experimental
REAPDOR fractions, AS/S
0
, as a function of the evolution times (T
r
= 0.1 ms). (Right) Curve fits for the
REAPDOR fractions at short dephasing (AS/S
0
s 0.3), using the function AS/S
0
= at
2
, corresponding to equation
(2.21).
Chapter 5. Framework structures of aluminosilicate binders 113
Finally, in order to assign the
29
Si resonance at o(
29
Si) = 83.2 ppm, the alkali-activated
metakaoline sample containing Si/Al = 1.0 and Na/Al = 1.0 has been studied by
29
Si{
27
Al}
REAPDOR. It is evident from the deconvolution of its
29
Si MAS NMR spectrum, Figure 5.6,
that the sample contains six non-equivalent
29
Si environments. However, the resonances at about
79 and 97 ppm, Figure 5.6 (e), have been neglected in the deconvolution of the REAPDOR
spectra due to their rather low intensity. Furthermore, the dipolar dephasing for the resonances at
87.5 ppm and 85.0 ppm were considered together. This assumption is reliable since both
resonances are derived from the Q
4
(4Al) components and have shown to have similar
magnitudes of dipolar dephasing, Table 5.1. The kn
Al
values for the resonances obtained from a
fit of their REAPDOR fractions are summarized in Table 5.1. It shows clearly that the resonance
at o(
29
Si) = 83.2 ppm possesses a dipolar dephase and an isotropic chemical shift corresponding
to a Q
3
(2Al) component.


Figure 5.6 Experimental (a) and simulated (b)
29
Si MAS NMR spectra of an alkali-activated metakaolin sample
(Si/Al = 1.0, Na/Al = 1.0). The experimental spectrum (a) was recorded at 7.05 T using a spinning speed of v
R
= 7.0
kHz, a 30-s relaxation delay and 2048 scans. The deconvolution of the spectrum reveals six different
29
Si
resonances, which are shown in (c), (d) and (e).
Chapter 5. Framework structures of aluminosilicate binders 114
As it is apparent from the data in Table 5.2, the eight
29
Si resonances observed in Figure
5.1 can be assigned unambiguously by considering their o(
29
Si) values together with the kn
Al

values obtained from the
29
Si{
27
Al} REAPDOR experiments. So far, it has not been possible to
assign the
29
Si resonance at o(
29
Si) = 78.6 ppm owing to its very low
29
Si intensity in the
studied sample. However, its isotropic chemical shift implies that it may originate from a non-
fully condensed tetrahedrally coordinated silicon environment.

Table 5.1 Isotropic chemical shifts and corresponding kn
Al
values for the eight
29
Si resonances observed for alkali-
activated metakaolin samples. The data are obtained from curve fits of the REAPDOR fractions at short dipolar
dephase (AS/S
0
s 0.3), using the function AS/S
0
= at
2
, corresponding to equation (2.21).


29
Si isotropic chemical shift in ppm
102.5 97.2

92.0 87.5 85.0 83.2
Figure 5.2(a) 0.39 0.51 0.25
Figure 5.3(c) 0.16 0.31 0.43 0.73
Figure 5.3(d) 0.19 0.29 0.44 0.67 0.69

Table 5.2 An assignment of the eight
29
Si resonances observed for the alkali-activated metakaolin samples. The
assignment is made in accordance with the results achieved from the
29
Si{
27
Al} REAPDOR experiments.


29
Si isotropic chemical shift in ppm
-107.1 -102.5 -97.2 -92.0 -87.5 -85.0 -83.2
Figure 5.2(a) - - - Q
4
(3Al) Q
4
(4Al) Q
4
(4Al) Q
3
(2Al)
Figure 5.3(c) Q
4
(0Al) Q
4
(1Al) Q
4
(2Al) Q
4
(3Al) Q
4
(4Al) - -
Figure 5.3(d) Q
4
(0Al) Q
4
(1Al) Q
4
(2Al) Q
4
(3Al) Q
4
(4Al) Q
4
(4Al) -

Chapter 5. Framework structures of aluminosilicate binders 115
5.2. Disorder in the double tetrahedral layer structure of Strtlingite
Strtlingite is a potential hydration product of aluminate-rich cements
[195]
. More recently,
this mineral has also been identified in alkali-activated blended cements, in which alkali
hydroxides are used to activate the hydration of supplementary cementitious materials
[196,197]
.
Strtlingite exhibits the ideal composition 2CaOAl
2
O
3
SiO
2
8H
2
O and crystallizes in the trigonal
space group R3m. A previous single-crystal X-Ray diffraction (XRD) study
[198]
of a mineral
sample revealed that the strtlingite structure consists of an octahedral brucite-type layer
[Ca
2
Al(OH)
6
2H
2
O]
+
and a double tetrahedral layer [(T,)
4
(OH,O)
8
0.25H
2
O]

. The octahedral
sites are fully occupied by Al
3+
while approx. 45% of the tetrahedral sites are vacant (). The
remaining 55 % are occupied by either Si
4+
or Al
3+
with an overall Si/Al molar ratio of 1:1.
Supplementary information about the double tetrahedral layer of strtlingite was provided by
another study
[199]
using
27
Al and
29
Si MAS NMR. That study identified four different tetrahedral
29
Si resonances from the
29
Si MAS NMR experiments (7.1 T), indicating four different SiO
4

environments of the double tetrahedral layer. Based on their isotropic chemical shifts, the three
almost equally intense
29
Si resonances in the region from 88.0 ppm to 81.0 ppm were assigned
to Q
2
(0Al), Q
2
(1Al) and Q
2
(2Al) components and a low intensity resonance at 110 ppm was
ascribed to the Q
4
(0Al) unit. This assignment is, however, in contrast to the basic structure of
strtlingite determined by XRD, which suggests that Al
3+
and Si
4+
occur mainly as Q
3
units. In
the same NMR study, a single tetrahedral AlO
4
resonance from the double tetrahedral layer and
an AlO
6
resonance from the brucite-type layer were observed. The resonances occur at 60.4 and
8.4 ppm in the
27
Al MAS NMR spectrum (7.05 T), respectively.
In this work the local structure of silicon and aluminum of the octahedral brucite-type
layer and the double tetrahedral layer have been reexamined using single-pulse
29
Si and
27
Al
MAS,
29
Si{
1
H} and
27
Al{
1
H} CP/MAS,
27
Al 3QMAS and
29
Si{
27
Al} REAPDOR NMR
experiments. However, this section only presents the results from the
29
Si{
27
Al} REAPDOR
NMR experiments for a synthetic sample of strtlingite, from which the
29
Si resonances are
assigned. In order to enhance the
29
Si sensitivity, a
29
Si{
1
H} CP/MAS period was applied prior
to the
29
Si{
27
Al} REAPDOR sequence as demonstrated by Figure 2.11 in Section 2.2.9. A
detailed discussion of the disordered layer structure of strtlingite is provided in manuscript 2.

Chapter 5. Framework structures of aluminosilicate binders 116
5.2.1.
29
Si MAS NMR
The
29
Si resonances of strtlingite shown in Figure 5.7 appear in the region from about
75 to 95 ppm, i.e. the isotropic chemical shift region for tetrahedral SiO
4
environments.
However, the overall lineshape of the spectrum differs somewhat from that presented in the
previous study of strtlingite. A
29
Si{
1
H} CP/MAS NMR experiment of the same sample clearly
resolves the lineshape into four well-separated
29
Si resonances with isotropic chemical shifts at
o(
29
Si) = 90.8 ppm, 86.4 ppm, 82.1 ppm and 79.7 ppm. Furthermore, a comparison of
Figure 5 (a) and (b) reveals an additional resonance at 84.9 ppm, which has to be included to
obtain satisfactory simulations of the spectra. As discussed above, the assignment of these
resonances from their isotropic chemical shifts alone appears to be uncertain, since the chemical
shift region from 75 ppm to 95 ppm may include any of the Q
2
(nAl) and Q
3
(nAl) structural
units. Furthermore, fully condensed Q
4
sites with a high number of Al incorporated in the
second-coordination sphere such as Q
4
(4Al) and Q
4
(3Al) may also exhibit isotropic chemical
shifts within this spectral region.


Figure 5.7
29
Si MAS (a) and
29
Si{
1
H} CP/MAS (b) NMR spectra (9.4 T) of a synthetic sample of alkali-containing
strtlingite. The spectra were recorded using spinning speeds v
R
of 6.0 and 3.0 kHz, respectively. Furthermore, the
29
Si{
1
H} CP/MAS NMR experiment was employed a contact time of 5.0 ms

Chapter 5. Framework structures of aluminosilicate binders 117
5.2.2.
29
Si{
27
Al} REAPDOR NMR
The assignment of the
29
Si resonances from the double tetrahedral layer of strtlingite has
been facilitated by supplementary information obtained using
29
Si{
27
Al} REAPDOR NMR
experiments. The ability of the REAPDOR experiment to retain SiAl dipolar couplings is
utilized to determine the number of Al atoms in the second-coordination sphere for each of the
29
Si resonances, as demonstrated in the previous sections. A plot of the REAPDOR fractions as a
function of evolution time for a synthetic sample of strtlingite is shown in Figure 5.8. The plot
clearly demonstrates that all
29
Si resonances are affected by the re-introduced SiAl dipolar
couplings. The corresponding kn
Al
values are determined from curve fits of REAPDOR
fractions to equation (2.21) and are summarized in Table 5.3. According to their isotropic
chemical shifts together with the kn
Al
values, the five
29
Si resonances representing five different
SiOAl environments of the double tetrahedral layer of strtlingite can be assigned to Q
3
(1Al):
90.8 ppm, Q
3
(2Al): 86.4 ppm, Q
3
(3Al): 83.8 ppm, Q
2
(1Al): 82.1 ppm and Q
2
(2Al): 79.7
ppm. This assignment is consistent with the fact that the resonances at 79.7 and 82.1 ppm
experience a more efficient
29
Si{
1
H} cross-polarization transfer than the remaining resonances
(cf. Figure 5.7), since they are linked to a larger number of OH groups.


Figure 5.8
29
Si{
27
Al} REAPDOR curves for a synthetic sample of strtlingite: () 90.8 ppm, (-) 86.4 ppm, ()
83.8 ppm, () 82.1ppm and () 79.7 ppm. (Left) Experimental REAPDOR fractions, AS/S
0
, as a function of
evolution time (T
r
= 0.1 ms). (Right) Curve fits for the REAPDOR fractions at short dephasing (AS/S
0
s 0.3), using
the function AS/S
0
= at
2
, corresponding to equation (2.21).

Chapter 5. Framework structures of aluminosilicate binders 118

Figure 5.9 Experimental (a) and simulated (b)
29
Si MAS NMR spectra of a synthetic sample of strtlingite. The
experimental spectrum (a) was recorded at 9.4 T using a spinning speed of v
R
= 6.0 kHz, a 30-s relaxation delay and
2048 scans. The deconvolution of the spectrum reveals five different resonances, which are shown in (c) and (d)
with isotropic chemical shifts at o(
29
Si) = 90.8 ppm, 86.4 ppm, 83.8 ppm, 82.1 ppm and 79.7 ppm.

Table 2.3 Isotropic chemical shifts, kn
Al
values determined from
29
Si{
27
Al} REAPDOR experiments and relative
intensities achieved from a deconvolution of the single-pulse
29
Si MAS NMR spectrum for a synthetic sample of
strtlingite.

Site assignment o(29Si) knAl Normalized intensity
29
Si: Q
3
(1Al) 90.8 ppm 0.21 2.8 %
29
Si: Q
3
(2Al) 86.4 ppm 0.44 55.2 %
29
Si: Q
3
(3Al) 83.8 ppm 0.65 10.3 %
29
Si: Q
2
(1Al) 82.1 ppm 0.25 22.7 %
29
Si: Q
2
(2Al) 79.7 ppm 0.48 8.9 %

Chapter 5. Framework structures of aluminosilicate binders 119
The distribution of silicon on the five different SiO
4
sites, reflected by their intensities in
the
29
Si MAS NMR spectrum, is obtained from a deconvolution, Figure 5.9; their normalized
intensities relative to the total intensity of the spectrum are Q
3
(1Al): 2.8 % , Q
3
(2Al): 55.2 %,
Q
3
(3Al): 10.3 %, Q
2
(1Al): 22.7 % and Q
2
(2Al) : 8.9 %. The result clearly demonstrates that the
double tetrahedral layer consists largely of Q
3
components, where Q represents either Si or Al in
tetrahedral coordination with three QOQ and one QOH bonds, forming a 2-dimensional
layer structure, which is consistent with the layer structure of 6-membered rings for strtlingite
reported from XRD
[198]
. Furthermore, this alumino-silicate network is disrupted by site vacancies
leading to approx. 30 % Q
2
units.

5.3. Summary
It has been demonstrated that the double-resonance REAPDOR sequence represents a
valuable tool in structural characterisation of cementitious materials. The
29
Si{
27
Al} REAPDOR
experiment has been applied to estimate the number of Al atoms incorporated in the second-
coordination sphere for SiO
4
tetrahedra and it has been shown that it provides a clear
discrimination between different Q
i
(nAl) environments. Unambiguous assignment of the
29
Si
resonances has been achieved by this approach for alkali-activated metakaolin samples and for
strtlingite.

Conclusions 120
Conclusion

In this project, the effects of fluoride ions on the formation of Portland clinker and its
hydration have been investigated. The modified clinkers were prepared using a white Portland
clinker from Aalborg Portland A/S as a main source. It is found that the mineralizing effect of
fluoride on the clinker formation associates with a coupled substitution Si
4+
+ O
2
Al
3+
+ F

in
the alite phase, promoting the formation of this phase by two different mechanisms. The first is a
thermodynamical effect, where the coupled incorporation of F

and Al
3+
stabilizes local regions
with the chemical composition Ca
27
Si
9-x
Al
x
(O
b
)
36
(O
i
)
9-x
(F
i
)
x
in the alite structure, where i and b
denote the interstitial and covalently bonded oxygen atoms, respectively. The second effect
arises from the mass balance of the coupled substitution, resulting in formal increase in the
quantity of belite which facilitates the reaction, CaO + belite alite. For the studied fluoride
contents (i.e. 0.04 0.77 %w), the mineralizing effects are increased with increasing fluoride
quantity.
For the hydration of fluoride-mineralized Portland cement, the existence of a critical bulk
fluoride content of about 0.36 %w F has been demonstrated. Below this value, an increase in the
fluoride content tends to increase the degree of hydration for the alite phase. However, it is
found that above the critical value, fluoride substantially affects the early hydration (i.e. the
hydration up to 28 days) for all cement phases by two tentative mechanisms. Firstly, the retarded
setting during the first 24 hours after mixing seems to be an intrinsic property of the fluoride-
mineralized alite; its hydrational activity during this period is very small. Secondly, at 24 hours
after mixing, the precipitation of the rather insoluble CaF
2
salt has been identified for the
fluoride-mineralized cements, where the CaF
2
quantity is increased with an increased fluoride
content in the anhydrous cements. This observation supports the model where the insoluble CaF
2

salt precipitates on the cement grain surface. At sufficiently high concentrations (i.e. for fluoride
contents above the critical value), a protective layer of CaF
2
may be formed, which retards the
cement hydration. However, this effect seems to be leveled out after 28 days of hydration.
From the optimizations of the Al
2
O
3
, Fe
2
O
3
and fluoride contents in Portland clinker, it is
found that the alite content can be increased by about 8 %w without additional CaO, which
implies that the CO
2
emission from the chemical reactions of the clinker formation is remained
at the same level as that of the unmodified white Portland clinker. Furthermore, hydration
experiments of the cement produced from this modified clinker show an increase by 36 % in its
Conclusions 121
one-day compressive strength as compared to the unmodified white Portland cement. For longer
hydration times, an increase of about 10 % in the compressive strength is observed for the
modified cement. The strength performance of blended cements produced from the modified
clinker using a 30 % replacement by different supplementary cementitious materials has also
been investigated. A remarkable reduction in the one-day compressive strength relative to that of
wPc is observed for all blended cements. However, for the long-term hydration (i.e. after 28 days
of hydration), the blended cement containing 70 % modified clinker, 10 % limestone filler and
20 % glass particles (developed in another PhD project of FUTURECEM) shows a reduction in
strength of only 5 % as compared to the white Portland cement.
Solid-state NMR has proven to be a valuable tool for structural investigation of the guest
ions (such as F

, Al
3+
and Fe
3+
) in anhydrous as well as hydrated phases of Portland cement. The
high sensitivity of the
19
F spins makes it possible to study the local structure environments of the
fluoride ions and their hydration processes in Portland cement. Moreover, it has been
demonstrated in this project that the couplings of fluoride to other guest ions may be investigated
using a combination of different advanced solid-state NMR techniques. In particular, the site
preferences for F

and Al
3+
guest ions in the calcium silicate phases of Portland clinker have
been investigated using
27
Al{
19
F} and
29
Si{
19
F} CP/MAS, and
29
Si{
19
F} CP-REDOR
experiments. It is found for the studied fluoride contents (> 0.77 %w F) that fluoride ions
substitute for the interstitial oxygen sites of the alite structure only. Furthermore, the
incorporation of fluoride ions in the alite structure is charge-balanced by a Si
4+
Al
3+

substitution on the tetrahedral site in the near vicinity of F

. This observation unambiguously


reveals the coupled substitution mechanism for the incorporation of F

and Al
3+
ions in the alite
structure, i.e. Si
4+
+ O
2
Al
3+
+ F

.
A new solid-state NMR pulse scheme (Forth and Back Cross Polarization), which is a
modification of the Cross-Polarization experiment, has been developed for studying the site
preferences of F

ions in the calcium silicate hydrate (CSH) phases of Portland cement. This
experiment utilizes the dipolar couplings to transfer the magnetization from the
19
F spins to
29
Si
spins and subsequently back to the
19
F spins, making it possible to selectively detect
19
F
resonances from
29
Si
19
F connectivities in the CSH structure within reasonable spectrometer
times. In this study, it is observed that a rather large amount of fluoride ions may be incorporated
in the CSH structure. The fluoride ions are either incorporated in the principal layer of the
jennite-like part or distributed in the interlayer of the tobermorite-like part of the CSH
structure. The incorporation of fluoride ions in the CSH structure tends to increase the average
silicate chain length for this phase.
Conclusions 122
The incorporation of the paramagnetic Fe
3+
ions in the calcium silicate phases of Portland
clinker has been investigated using the
29
Si Inversion-Recovery NMR experiment and X-ray
powder diffraction. These experiments show clear evidence for the presence of paramagnetic
Fe
3+
ions in the alite as well as the belite structures. It is also demonstrated that the Fe
3+
ions tend
to be incorporated in the octahedral sites of alite by substitution for the Ca
2+
ions.
Finally, the applicability of the
29
Si{
27
Al} REAPDOR experiment in structural
characterizations of the aluminosilicate networks structure has been demonstrated in this project
for a series of alkali-activated metakaoline samples and a synthetic sample of strtlingite. This
NMR experiment allows a determination of the number of Al atoms incorporated in the second-
coordination sphere for the probed silicon atoms and thereby, discriminating between different
SiOAl connectivities.

References 123
References
[1] K. Salazar and M.K. McNutt. Mineral commodity summaries 2011. U.S. Department of
the Interior and U.S. Geological Survey. United States Goverment Printing Office,
Washington, 2011.
[2] Cement industry energy and CO
2
performance "Getting the Number Right". Prepared by
the World Business Council for Sustainable Development - The Cement Sustainability
Initiative, 2009.
[3] E. Gartner. Industrially interesting approaches to low-CO
2
cements. Cem. Concr. Res.,
2004, 34, 1489-1498.
[4] J.S. Damtoft, J. Lukasik, D. Herfort, D. Sorrentino and E.M. Gartner. Sustainable
development and climate change initiatives. Cem. Concr. Res., 2008, 38, 115-127.
[5] Global Cement Database on CO
2
and Energy Information. Prepared by the World
Business Council for Sustainable Development - Cement Sustainability Initiative.
www.wbcsdcement.org/co2data.
[6] Cement Technology Roadmap 2009 - Carbon emissions reductions up to 2050. Prepared
by the World Business Council for Sustainable Development and the International Energy
Agency. https://www.wbcsd.org.
[7] M. Schneider, M. Romer, M. Tschudin and H. Bolio. Sustainable cement production-
present and future. Cem. Concr. Res., 2011, 41, 642-650.
[8] C. Shi, A.F. Jimnez and A. Palomo. New cement for the 21
st
century: The pursuit of an
alternative to Portland cement. Cem. Concr. Res., 2011, 41, 750-763.
[9] E.G. Shame and F.P. Glasser. Stable Ca
3
SiO
5
Solid Solutions containing fluorine and
aluminium made between 1050 and 1250
o
C. Br. Ceram. Trans. J., 1987, 86, 13-17.
[10] G.K. Moir. Improvements in the early strength properties of Portland cement. Phil.
Trans. R. Soc. Lond. A, 1983, 310, 127-138.
[11] J. Aspdin. Producing an artificial stone. 1824, British Patent BP 5022.
References 124
[12] The European Stardard EN 197-1, 2000. Composition, specification and conformity
criteria for common cements.
[13] H.F.W. Taylor. Cement chemistry. 2
nd
ed, 1997, Thomas Telford Publishing.
[14] V. Zetola. Differences in basic standards of common cements. Cem. Concr. Res., 2011,
41, 764-766.
[15] A.K. Chatterjee. Chemistry and engineering of the clinkerization process - Incremental
advances and lack of breakthroughs. Cem. Concr. Res., 2011, 41, 624-641.
[16] F. Sorrentino. Chemistry and engineering of the production process: State of the art.
Cem. Concr. Res., 2011, 41, 616-623.
[17] J.I. Bhatty, F.M. Miller, S.H. Kosmatka and R.P. Bohan (eds.). Innovations in Portland
cement manufacturing. Portland Cement Association in Skokie, 2004, Illinois, USA.
[18] A.A. Phillip, C. Hung, T. Herman. The cement plant operating handbook. 5
th
ed. ICR,
Tradeship Publications Ltd, 2007, Surrey, UK.
[19] A. Wolter. Influence of the kiln system on the clinker properties. Zement-Kalk-Gips,
1985, 38 (10), 612.
[20] N.I. Golovastikov, R.G. Matveeva and N.V. Belov. Crystal structure of the tricalcium
silicate. Kristallografiya, 1975, 20, 721-729.
[21] F. Nishi, Y. Takeuchi and I. Maki. Tricalcium silicate: The monoclinic superstructure. Z.
Kristallogr., 1985, 172, 297-314.
[22] A.M. Il'inets, Y.A. Malinovskii and N.N. Nevskii. The crystal structure of the
rhombohedral modification of tricalcium silicate. Doklady Akademii Nauk SSSR, 1985,
281, 332-336.
[23] M. Bigar, A. Guinier, C. Mazires, M. Regourd, N. Yannaquis, W. Eysbl. Polymorphism
of tricalcium silicate and its solid solutions. J. Am. Ceram. Soc., 1967, 50, 609-619.
References 125
[24] G.K. Moir and F.P. Glasser. Mineralizers, modifiers and activators in the clinkering
process. Proc. 9
th
International Congress on the Chemistry of Cement, New Delhi, India,
1992, Vol. I, 125-152.
[25] V. Johansen and Bhatty J. I. Fluxes and mineralizers in clinkering process. In: J.I. Bhatty,
F.M. Miller, and S.H. Kosmatka (eds.), Innovations in Portland cement manufacturing,
2004, Section 3.5, 369-402.
[26] I. Maki and K. Goto. Factors influencing the phase constitution of alite in portland
cement clinker. Cem. Concr. Res., 1982, 12, 301-308.
[27] A.B. Diouri A., J. Aride, F. Puertas and T. Vazquez. Stable Ca
3
SiO
5
solid solution
containg Manganese and Phousporus. Cem. Concr. Res., 1997, 27, 1203-1212.
[28] W. Sinclair and G.W. Groves. Transmission Electron Microscopy and X-Ray Diffraction
of doped tricalcium silicate. J. Am. Ceram. Soc., 1984, 61, 325-330.
[29] A.M. Harrison, H.F.W. Taylor and N.B. Winter. Electron-Optical analyses of the phases
in a Portland cement clinker, with some observations on the calculation of quantitative
phase composisition. Cem. Concr. Res., 1985, 15, 775-780.
[30] A. Ghose and P. Barnes. Electron Microprobe Ananlysis of Portland cement clinker.
Cem. Concr. Res., 1979, 9, 747-755.
[31] M. Kristmann. Portland cement clinker mineralogi and chemical investigations. Cem.
Concr. Res., 1978, 8, 93-102.
[32] I. Nettleship, K.G. Slavick, Y.J. Kim and W.M. Kriven. Phase transformation in
dicalcium silicate: I, Fabrication of phase stability of fine-grained | phase. J. Am.
Ceram. Soc., 1992, 75, 2400-2406.
[33] K.H. Jost, B. Ziemer and R. Seydel. Redetermination of the structure of |-dicalcium
silicate. Acta Cryst., 1977, B33, 1696-1700.
[34] P. Mondal and J.W. Jeffery. The crystal structure of tricalcium aluminate, Ca
3
Al
2
O
6
.
Acta Cryst., 1975, B31, 689-697.
References 126
[35] L. Gobbo, L. Sant Agostino and L. Garcez. C
3
A polymorphs related to industrial clinker
alkalies content. Cem. Concr. Res., 2004, 34, 657-664.
[36] I. Maki. Morphology of the so-called prismatic phase in Portland cement clinker. Cem.
Concr. Res., 1974, 4, 87-97.
[37] F. Puertas, M.T. Blanco Varela and R. Dominguez. Characterization of Ca
2
AlMnO
5
. A
comparative study between Ca
2
AlMnO
5
and Ca
2
AlFeO
5
. Cem. Concr. Res., 1990, 20,
429-438.
[38] L. Bonafous, C. Bessada, D. Massiot and J.P. Coutures.
29
Si MAS NMR study of
dicalciums silicate: The structural influence of sulfate and alumina stabilizers. J. Am.
Ceram. Soc., 1995, 78, 2603-2608.
[39] S.L. Poulsen, H.J. Jakobsen and J. Skibsted. Incorporation of phosphorus guest ions in
the calcium silicate phases of Portland cement from
31
P MAS NMR spectroscopy. Inorg.
Chem., 2010, 49, 5522-5529.
[40] M.N. Noirfontaine, S. Tusseau-Nenez, M. Signes-Frehel, G. Gasecki and C. Girod-
Labiance. Effect of phosphorus impurity on tricalcium silicate T
1
: From synthesis to
structural characterization. J. Am. Ceram. Soc., 2009, 92, 2337-2344.
[41] H.E. Borgholm, D. Herfort and S. Rasmussen. A new blended cement based on
mineralised clinker. World cement reseach and development, 1995, August, 27-33.
[42] M.T. Blanco-Varela, A. Palomo, F. Puertas and T. Vzquez. CaF
2
and CaSO
4
in white
cement clinker production. Adv. Cem. Res., 1997, 35, 105-113.
[43] W.A. Klemm, I. Jawed and K.J. Holub. Effects of calcium fluoride mineralization on
silicates and melt formation in portland cement clinker. Cem. Concr. Res., 1979, 9, 489-
496.
[44] W.A. Klemm and I. Jawed. Mineralizers and fluxes in the clinkering process. Proc. 9
th

International Congress on the Chemistry of Cement, New Delhi, India, 1992, Vol. II, 150-
155.
References 127
[45] Y.B. Pliego-Cuervo and F.P. Glasser. The role of sulphate in cement clinkering:
subsolidus phase relations in the system CaO-Al
2
O
3
-SiO
2
-SO
3
. Cem. Concr. Res., 1979,
9, 51-56.
[46] J. Skibsted, M.D. Andersen and H.J. Jakobsen. Structural investigations of Portland
cement components, hydration, and effects of admixtures by solid-state NMR
spectroscopy. Proc. 16
th
Internationale Baustofftagung, IBAUSIL conference, Weimer,
2006, Vol. I, 0297-0313.
[47] J. Ambroise, S. Maximilien and J. Pera. Properties of metakaolin blended cements. Adv.
Cem. Based Mater., 1994, 1, 161-168.
[48] H.F.W. Taylor, K. Mohan and G.K. Moir. Ananlytical study of pure extended Portland
cement pastes: II, Fly ash- and slag-cement pastes. J. Am. Ceram. Soc., 1985, 68, 685-
690.
[49] J. Hjorth, J. Skibsted and H.J. Jakobsen.
29
Si MAS NMR studies of Portland cement
components and effects of microsilica on the hydration reaction. Cem. Concr. Res., 1988,
18, 789-798.
[50] B. Lothenbach, G. Le Saout, E. Gallucci and K. Scrivener. Influence of limestone on the
hydration of Portland cements. Cem. Concr. Res., 2008, 38, 848-860.
[51] J. Bensted and P. Barnes (eds.). Structure and Performance of Cements. 2nd ed., 2002,
Spon Press, New York., p. 57-113.
[52] H.M. Jennings, B.J. Dalgleish and P.L. Pratt. Morphological development of hydrating
tricalcium silicate as examined by Electron Microscopy techniques. J. Am. Ceram. Soc.,
1981, 64, 567-572.
[53] E. Breval. C
3
A Hydration. Cem. Concr. Res., 1976, 6, 129-138.
[54] M. Collepardi, G. Baldini and M. Pauri. Tricalcium aluminate hydration in the presence
of lime, gypsum or sodium sulfate. Cem. Concr. Res., 1978, 8, 571-580.
[55] E. Gallucci, P. Mathur and K. Scrivener. Microstructural development of early age
hydration shells around cement grains. Cem. Concr. Res., 2010, 40, 4-13.
References 128
[56] G.W. Groves. Portland cement clinker viewed by transmission electron microscopy. J.
Mater. Sci., 1981, 16, 1063-1070.
[57] I.G. Richardson. The nature of C-S-H in hardened cements. Cem. Concr. Res., 1999, 29,
1131-1147.
[58] L.S.D. Glasser, E.E. Lachowski, K. Mohan and H.F.W. Taylor. A multi-method study of
C
3
S hydration. Cem. Concr. Res., 1978, 8, 733-740.
[59] H.F.W. Taylor and D.E. Newbury. Calcium hydroxide distribution and calsium silicate
hydrate composition in tricalcium silicate and |-dicalcium silicate pastes. Cem. Concr.
Res., 1984, 14, 93-98.
[60] J.D. Bernal, J.W. Jeffery and H.F.W. Taylor. Crystallographic research on the hydration
of Portland cement. A first report on investigations in progress. Mag. Concr. Res., 1952,
4, 49-54.
[61] H.F.W. Taylor. Proposed structure for calcium silicate hydrate gel. J. Am. Ceram. Soc.,
1986, 69, 464-467.
[62] I.G. Richardson and G.W. Groves. Models for the composition and structure of calcium
silicate hydrate (C-S-H) gel in hardened tricalcium silicate pastes. Cem. Concr. Res.,
1992, 22, 1001-1010.
[63] X. Cong and R.J. Kirkpatrick.
29
Si MAS NMR study of the structure of calcium silicate
hydrate. Adv. Cem. Based Mater., 1996, 3, 144-156.
[64] I.G. Richardson. The calcium silicate hydrates. Cem. Concr. Res., 2008, 38, 137-158.
[65] I.G. Richardson and G.W. Groves. The incorporation of minor and trace elements into
calcium silicate hydrate (C-S-H) gel in hardened cement pastes. Cem. Concr. Res., 1993,
23, 131-138.
[66] E. Bonaccorsi and S. Merlino. The crystal structure of Tobermorite 14 (Plombierite), a
C-S-H phase. J. Am. Ceram. Soc., 2005, 88, 505-512.
References 129
[67] E. Bonaccorsi, S. Merlino and H.F.W. Taylor. The crystal structure of Jennite,
Ca
9
Si
6
O
18
(OH)
6
8H
2
O. Cem. Concr. Res., 2004, 34, 1481-1488.
[68] D.M. Henderson and H.S. Gutowsky. A nuclear magnetic resonance determination of the
hydrogen positions in Ca(OH)
2
. American Mineralogist, 1962, 47, 1231-1251.
[69] R.R. Ernst, G. Bodenhausen and A. Wokaun. Principles of Nuclear Magnetic Resonance
in one and two Dimensions. 1987, Clarendon Press, Oxford.
[70] J. Mason. Conventions for the reporting of nuclear magnetic shielding (or shift) tensors
suggested by participants in the NATO ARW on NMR Shielding Constants at the
University of Maryland, College Park, July 1992. Solid State Nucl. Magn. Res., 1993, 2,
285-288.
[71] M.J. Duer (eds.). Solid-State NMR Spectroscopy, principles and applications. Blackwell
Science Ltd, 2002, Great Britain.
[72] J. Herzfeld and A.E. Berger. Sideband intensities in NMR spectra of samples spinning at
the magic angle. J. Chem. Phys., 1980, 73, 6021-6030.
[73] M.R. Hansen, H.J. Jakobsen and J. Skibsted.
29
Si Chemical Shift Anisotropies in calcium
silicates from high-field
29
Si MAS NMR spectroscopy. Inorg. Chem., 2003, 7, 2368-2377.
[74] J. Skibsted, N.C. Nielsen, H. Bildse and H.J. Jakobsen. Satellite transitions in MAS
NMR spectra of quadrupolar nuclei. J. Magn. Res., 1991, 95, 88-117.
[75] M. Lee and W. Goldburg. Nuclear-Magnetic-Resonance line narrowing by a rotating rf
Field. Phys. Rev., 1965, 140, 1261-1271.
[76] A.E. Bennett, C.M. Rienstra, M. Auger and L.V. Lakshmi. Heteronuclear decoupling in
rotating solids. J. Chem. Phys, 1995, 103, 6951-6958.
[77] A.J. Shaka, J. Keeler, T. Frenkiel and R. Freeman. An improved sequence for broadband
decoupling: WALTZ-16. J. Magn. Res., 1983, 52, 335-338.
References 130
[78] S.L. Poulsen, V. Kocaba, G.L. Saout, H.J. Jakobsen, K.L. Scrivener and J. Skibsted.
Improved quantification of alite and belite in anhydrous Portland cements by
29
Si MAS
NMR: Effects of paramagnetic ions. Solid State Nucl. Magn. Res., 2009, 36, 32-44.
[79] E. Lippmaa, M. Mgi, A. Samoson and G. Engelhardt. Structural studies of silicates by
solid-state high-resolution
29
Si NMR. J. Am. Chem. Soc., 1980, 102, 4889-4893.
[80] J. Klinowski. Nuclear Magnetic Resonance studies of zeolites. Prog. NMR Spec., 1984,
16, 237-309.
[81] G. Engelhardt. Silicon-29 NMR of solid silicates. In: D.M. Grant and R.K. Harris (eds.),
Encyclopedia of Nuclear Magnetic Resonance, John Wiley & Sons, 1996, New York, p.
4398-4407.
[82] J. Skibsted and H.J. Jakobsen. Characterization of the calcium silicate and aluminate
phases in anhydrous and hydrated Portland cements by
27
Al and
29
Si MAS NMR
spectroscopy. In: P. Colombet, A.R. Grimmer, H. Zanni, P. Sozzani (eds.), Nuclear
Magnetic Resonance Spectroscopy of Cement-Based Materials. Springer-Verlag, 1998,
Berlin Heidelberg, p. 3-45.
[83] J. Skibsted, H.J. Jakobsen and C. Hall. Quantification of calcium silicate phases in
Portland cements by
29
Si MAS NMR spectroscopy. J. Chem. Soc. Faraday Trans., 1995,
91, 4423-4430.
[84] A.R. Grimmer, F. Von Lampe, M. Mgi and E. Lippmaa. High-resolution solids-state
29
Si NMR of polymorphs of Ca
2
SiO
4
. Cem. Concr. Res., 1985, 15, 467-473.
[85] J. Skibsted, J. Hjorth and H.J. Jakobsen. Correlation between
29
Si NMR chemical shifts
and mean Si---O bond lengths for calcium silicates. Chem. Phys. Lett., 1990, 172, 279-
283.
[86] J. Skibsted, E. Henderson and H.J. Jakobsen. Characterization of calcium aluminate
phases in cements by
27
Al MAS NMR spectroscopy. Inorg. Chem., 1993, 32, 1013-1027.
[87] A. Labouriau, Y.W. Kim, S. Chipera, D.L. Bish and W.L. Earl. A
19
F Nuclear Magnetic
Resonance study of natural clays. Clays and Clay Minerals, 1995, 43, 697-704.
References 131
[88] R.E. Youngman and S. Sen. Structural role of fluorine in amorphous silica. J. Non-Cryst.
Solids, 2004, 349, 10-15.
[89] H. Koller, A. Wlker, L.A. Villaescusa, M.J. Das-Cabanas and Valencia, S. and
Camblor, M. A. Five-coordnate silicon in high-silica zeolites. J. Am. Chem. Soc., 1999,
121, 3368-3376.
[90] Y. Liu and H. Nekvasil. Si-F bonding in aluminosilicate glasses: Inferences from ab
initio NMR calculations. American Mineralogist, 2002, 87, 339-346.
[91] J. Skibsted, M.D. Andersen and H.J. Jakobsen. Applications of solid-state Nuclear
Magnetic Resonance (NMR) in studies of Portland cement-based materials. Zement-
Kalk- Gips, 2007, 60 (6), 77-83.
[92] K.J.D. Mackenzie and M.E. Smith. Multinuclear Solid-State NMR of inorganic materials.
Pergamon, Elsevier science Ltd, 2002, Oxford, UK, p. 550-563.
[93] I.L. Moudrakovski, R. Alizadeh and J.J. Beaudoin. Natural abundance high field
43
Ca
solid state NMR in cement science. Phys. Chem. Chem. Phys., 2010, 12, 6961-6969.
[94] G.M. Bowers and R.J. Kirkpatrick. Natural abundance
43
Ca NMR spectroscopy of
Tobermorite and Jennite: Model compounds for C-S-H. J. Am. Ceram. Soc., 2009, 92,
545-548.
[95] Z. Lin and M.E. Smith. Probing the local structural environment of calcium by natural-
abundance solid-state
43
Ca NMR. Phys. Rev. B, 2004, 69, 224107-1-224107-7.
[96] K.J.D. Mackenzie, M.E. Smith and A. Wong. A multinuclear MAS NMR study of
calcium-containing aluminosilicate inorganic polymers. J. Mater. Chem., 2007, 17, 5090-
5096.
[97] F. Angeli, M. Gaillard, P. Jollivet and T. Charpentier. Contribution of
43
Ca MAS NMR for
probing the structural configuration of calcium in glass. Chem. Phys. Lett., 2007, 440,
324-328.
References 132
[98] A. Wong, D. Laurencin, R. Dupree and M.E. Smith. Two-dimensional
43
Ca-
1
H
correlation solid-state NMR spectroscopy. Solids State Nucl. Magn. Res., 2009, 35, 32-
36.
[99] K. Shimoda, Y. Tobu, Y. Shimoikeda, T. Nemoto and K. Saito. Muliple Ca
2+

environments in silicate glasses by high-resolution
43
Ca MQMAS NMR technique at high
and ultra-high (21.8 T) magnetic fields. J. Magn. Res., 2007, 186, 156-159.
[100] D. Laurencin, C. Gervais, A. Wong, C. Coelho, F. Mauri and D. Massiot. Implementation
of High Resolution
43
Ca Solid State NMR Spectroscopy: Toward the elucidation of
calcium sites in biological materials. J. Am. Chem. Soc., 2009, 131, 13430-13440.
[101] I.C.M. Kwan, A. Wong, Y.M. She, M.E. Smith, and G. Wu. Direct NMR evidence for
Ca
2+
ion binding to G-quartets. Chem. Commun., 2008, 682-684.
[102] R.L. Vold, J.S. Waugh, M.P. Klein and D.E. Phelps. Measurement of spin relaxation in
complex systems. J. Chem. Phys., 1968, 48, 3831-3832.
[103] M.R. McHenry and B.G. Silbernagel. Nuclear Spin-Lattice relaxation in the La
1-c
Gd
c
Al
2

intermetallic compounds. Phys. Rev. B, 1972, 5, 2958-2972.
[104] S. Hayashi and E. Akiba. Nuclear spin-lattice relaxation mechanisms in kaolinite
confirmed by magic-angle spinning. Solid State Nucl. Magn. Res., 1995, 4, 331-340.
[105] S.R. Hartmann and E.L. Hahn. Nuclear double resonance in the rotating frame. Phys.
Rev., 1962, 128, 2042-2053.
[106] W. Kolodziejski and J. Klinowski. Kinetics of Cross-Polarization in solid-state NMR: A
Guide for Chemists. Chem. Rev., 2002, 102, 613-628.
[107] J. Skibsted, S. Rasmussen, D. Herfort and H.J. Jakobsen.
29
Si cross-polarization magic-
angle spinning NMR spectroscopy - an efficient tool for quantification of thaumasite in
cement-based materials. Cem. Concr. Compos., 2003, 25, 823-829.
[108] G. Metz, X.L. Wu and S.O. Smith. Ramped-Amplitude Cross Polarization in Magic-
Angle-Spinning NMR. J. Magn. Res. A, 1994, 110, 219-227.
References 133
[109] M. Sardashti and G.E. Maciel. Effects of sample spinning on Cross Polarization. J. Magn.
Res., 1987, 72, 467-474.
[110] B.H. Meier. Cross polarization under fast magic angle spinning: thermodynamical
consideration. Chem. Phys. Lett., 1992, 188, 201-207.
[111] A.J. Vega. CP/MAS of quadrupolar S = 3/2 nulcei. J. Magn. Res., 1992, 1, 17-32.
[112] A.J. Vega. MAS NMR spin locking of half-integer quadrupolar nuclei. J. Magn. Res.,
1992, 96, 50-68.
[113] J. Schaefer, E.O. Stejskal, J.R. Garbow and R.A. McKay. Quantitative determination of
the concentrations of
13
C-
15
N chemical bonds by Double Cross-Polarization NMR. J.
Magn. Res., 1984, 59, 150-156.
[114] M. Wilhelm, H. Feng, U. Tracht and H.W. Spiess. 2D CP/MAS
13
C isotropic chemical
shift correlation established by
1
H spin diffusion. J. Magn. Res., 1998, 134, 255-260.
[115] S. Saburi, A. Kawahara, C. Henmi, I. Kuschi and K. Kihara. The refinement of the crystal
structure of cuspidine. Mineral. J., 1977, 8, 286-298.
[116] A. Brinkmann and M.H. Levitt. Symmetry principles in the nuclear magnetic resonance
of spinning solids: Heteronuclear recoupling by generalized Hartmann-Hahn sequences.
J. Chem. Phys., 2001, 115, 357-383.
[117] T. Gullion and J. Schaefer. Rotational-Echo Double-Resonance NMR. J. Magn. Res.,
1989, 81, 196-200.
[118] B.V. Rossum, H. Frster and H.J.M.d. Groot. High-field and high-speed CP-MAS
13
C
NMR heteronuclear dipolar-correlation spectroscopy of solids with frequency-switched
Lee-Goldburg homonuclear decoupling. J. Magn. Res., 1997, 124, 516-519.
[119] T. Gullion. Introduction to Rotational-Echo, Double-Resonance NMR. Concepts Magn.
Reson., 1998, 10, 277-289.
[120] K.T. Mueller. Analytic solutions for the time evolution of dipolar-dephasing NMR
signals. J. Magn. Res. A, 1995, 113, 81-93.
References 134
[121] C.A. Fyfe and A.R. Lewis. Investigation of the viability of solid-state NMR distance
determination in multiple spin systems of unknown structure. J. Phys. Chem. B, 2000,
104, 48-55.
[122] T. Gullion and A.J. Vega. Measuring heteronuclear dipolar couplings for I = 1/2, S > 1/2
spin pairs by REDOR and REAPDOR NMR. Progress in NMR Spec., 2005, 47, 123-136.
[123] M. Bertmer, L. Zuchner, J.C.C. Chan and H. Eckert. Short and medium range order in
sodium aluminoborate glasses. 2. Site connectivities and cation distributions studied by
Rotational Echo Double Resonance NMR Spectroscopy. J. Phys. Chem. B, 2000, 104,
6541-6553.
[124] M.R. Hansen, H.J. Jakobsen and J. Skibsted. Structural environments for boron and
aluminum in alumina-boria catalysts and their precursors from
11
B and
27
Al Single- and
Double-Resonance MAS NMR Experiments. J. Phys. Chem. C, 2008, 112, 7210-7222.
[125] L. Zhang, C.C.d. Araujo and H. Eckert. Structural role of fluoride in aluminophosphate
sol-gel glasses: High-resolution Double-Resonance NMR studies. J. Phys. Chem. B,
2007, 111, 10402-40412.
[126] C.A. Fyfe, D.H. Brouwer, R.H. Levis and J.M. Chzeau. Location of the fluoride ion in
tetrapropylammonium fluoride silicalite-1 determined by H/F/Si triple resonance CP,
REDOR, and TEDOR NMR experiments. J. Am. Chem. Soc., 2001, 123, 6882-6891.
[127] A.A. Maudsley. Modified Carr-Purcell-Meiboom-Gill sequence for NMR fourier imaging
applications. J. Magn. Res., 1986, 69, 488-491.
[128] T. Gullion, D.B. Baker and M.S. Conradi. New, compensated Carr-Purcell sequences. J.
Magn. Res., 1990, 89, 479-484.
[129] T. Gullion. Measurement of dipolar interactions between spin-1/2 and quadrupolar
nuclei by rotational-echo, adiabatic-passage, double-resonance NMR. Chem. Phys. Lett.,
1995, 246, 325-330.
[130] E. Hughes, T. Gullion, A. Goldbourt, S. Vega and A.J. Vega. Internuclear distance
determination of S=1, I=1/2 spin pairs using REAPDOR NMR. J. Magn. Res., 2002, 156,
230-241.
References 135
[131] M. Bertmer and H. Eckert. Dephasing of spin echoes by multiple heteronuclear dipolar
interactions in rotational echo double resonance NMR experiments. Solid State Nucl.
Magn. Reson., 1999, 15, 139-152.
[132] L. Zuchner, J.C.C. Chan, W. Muller-Warmuth and H. Eckert. Short-Range Order and Site
Connectivities in Sodium Aluminoborate Glasses: I. Quantification of Local
Environments by High-Resolution
11
B,
23
Na, and
27
Al Solid-State NMR. J. Phys. Chem. B,
1998, 102, 4495-4506.
[133] R.K. Harris, E.D. Becker, S.M.C. de Menezes, R. Goodfellow and P. Granger. NMR
Nomenclature. Nuclear Spin Properties and Conventions for Chemical Shifts (IUPAC
Recommentdations 2001). Pure Appl. Chem., 2001, 73, 1795-1818.
[134] Q. Luo, J. Yang, W. Hu, M. Zhang, Y. Yue, C. Ye. Unambiguously distinguishing
Si[3Si,1Al] and Si[3Si,1OH] stuctural units in zeolite by
1
H/
29
Si/
27
Al triple-resonance
solid-state NMR spectroscopy. Solid State NMR, 2005, 28, 9-12.
[135] S. Ganapathy, R. Kumar, V. Montouillout, C. Fernandez and J.P. Amoureux.
Identification of tetrahedrally ordered SiOAl environments in molecular sieves by
{
27
Al}
29
Si REAPDOR NMR. Chem. Phys. Lett., 2004, 390, 79-83.
[136] J.B. Nicholas, R.E. Winans, R.J. Harrison, L.E. Iton, L.A. Curtiss and A.J. Hopfinger. Ab
initio molecular orbital study of the effects of basis set size on the calculated structure
and acidity of hydroxyl groups in framework molecular sieves. J. Phys. Chem., 1992, 96,
10247-10257.
[137] C.A. Fyfe, K.T. Mueller and G.T. Kokotailo. Solid-State NMR Studies of Zeolites and
Related Systems. In: A.T. Bell and A. Pines. NMR Techniques in Catalysis (eds.). Marcel
Dekker, Inc., 1994, New York.
[138] L. Frydman and J.S. Harwood. Isotropic spectra of half-integer quadrupolar spins from
bidimensional Magical-Angle Spinning NMR. J. Am. Chem. Soc., 1995, 117, 5367-5368.
[139] D. Massiot, B. Touzo, D. Trumeau, J.P. Coutures, J. Virlet, P. Florian. Two-dimensional
magic-angle spinning isotropic reconstruction sequences for quadrupolar nuclei. Solid
State NMR, 1996, 6, 73-83.
References 136
[140] J.P. Amoureux, C. Fernandez and L. Frydman. Optimized multiple-quantum magic-angle
spinning NMR experiments on half-integer quadrupoles. Chem. Phys. Lett., 1996, 259,
347-355.
[141] J. Amoureux, C. Fernandez and S. Steuernagel. Z-Filtering in MQMAS NMR. J. Magn.
Res. Ser. A, 1996, 123, 116-118.
[142] A. Goldbourt and S. Vega. Signal ehancement in 5QMAS spectra of spin-5/2 quadrupolar
Nuclei. J. Magn. Res., 2002, 154, 280-286.
[143] G. Wu, D. Rovnyak and R.G. Griffin. Quantitative multiple-quantum Magic-Angle-
Spinning NMR spectroscopy of quadrupolar nuclei in solids. J. Am. Chem. Soc., 1996,
118, 9326-9332.
[144] S.H. Wang, S.M.D. Paul and L.M. Bull. High-resolution heteronuclear correlation
between quadrupolar and spin- nuclei using Multiple-Quantum Magic-Angle Spinning.
J. Magn. Res., 1997, 125, 364-368.
[145] A. Medek, J.S. Harwood and L. Frydman. Multiple-Quantum Magic-Angle Spinning
NMR: A new method for the study of quadrupolar nuclei in solids. J. Am. Chem. Soc.,
1995, 117, 12779-12787.
[146] H. Saalfeld and M. Wedde. Refinement of the crystal structure of gibbsite, Al(OH)
3
. Z.
Kristallogr., 1974, 139, 129-135.
[147] S.E. Ashbrook, J. McManus, K.J.D. Mackenzie and S. Wimperis. Multiple-Quantum and
Cross-Polarized
27
Al MAS NMR of mechanically treated mixtures of kaolinite and
gibbsite. J. Phys. Chem. B, 2000, 104, 6408-6416.
[148] C. Gilioli, F. Massazza and M. Pezzuoli. Studies on clinker calcium silicates bearing
CaF
2
and CaSO
4
. Cem. Concr. Res., 1979, 9, 295-302.
[149] L. Kacimi, A. Simon-Masseron, A. Ghomari and Z. Derriche. Influence of NaF, KF, and
CaF
2
addition on the clinker burning temperature and its properties. C. R. Chimie, 2006,
9, 154-163.
References 137
[150] J. Mukerji. Phase equilibrium diagram CaO-CaF
2
-2CaOSiO
2
. J. Am. Ceram. Soc., 1965,
48, 210-213.
[151] S. Giminez-Molina, M.T. Blanco, T. Marr and F.P. Glasser. Phase relations in the system
Ca
2
SiO
4
-CaO-CaSO
4
-CaF
2
relevant to cement clinkering. Adv. Cem. Res., 1991, 14, 81-
86.
[152] J.K. Christie, A. Pedone, M.C. Menziani and A. Tilocca. Fluoride evironment in
bioactive glasses: ab initio molecular dynamics simulations. J. Phys. Chem., 2011, 115,
2038-2045.
[153] D.S. Brauer, N. Karpukhina, R.V. Law and R.G. Hill. Structure of fluoride-containing
bioactive glasses. J. Mater. Chem., 2009, 19, 5629-5636.
[154] I. Odler and S. Abdul-Maula. Structure and properties of Portland cement clinker doped
with CaF
2
. J. Am. Ceram. Soc., 1980, 63, 654.
[155] R.T.H. Aldous. The hydraulic behaviour of rhombohedral alite. Cem. Concr. Res., 1983,
13, 89-96.
[156] B. Jansang, A. Nonat and J. Skibsted. Modelling of guest-ion incorporation in the
anhydrous calcium silicate phases of Portland cement by periodic Density Functional
Theory calculations. Proc. CONMOD 2010, 3
rd
RILEM Conference on Concrete
Modelling, 2010, Lausanne, Switzerland. p. 25-28.
[157] A.R. West. Chapter 5: Crystal defects, non-stoichiometry and solid solution. In: Basic
Solid State Chemistry, 2
nd
ed. John Wiley & Sons Ltd. 2000.
[158] E.M. Gartner and F.J. Tang. Formation and properties of high sulfate Portland cement
clinkers. il cemento, 1987, 2, 141-164.
[159] A. Crumbie, G. Walenta and T. Fllmann. Where is the iron? Clinker microanalysis with
XRD Rietveld, optical microscopy/point counting, Bogue and SEM-EDS techniques. Cem.
Concr. Res., 2006, 36, 1542-1547.
[160] C.L. Edwards, L.B. Alemany and A.R. Barron. Solid-state
29
Si NMR analysis of cements:
comparing different methods of relaxation analysis for determining spin-lattice relaxation
References 138
times to enable determination of the C
3
S/C
2
S ratio. Ind. Eng. Chem. Res., 2007, 46, 5122-
5130.
[161] W.A. Gutteridge. On the dissolution of the interstitial phases in Portland cement. Cem.
Concr. Res., 1979, 9, 319-324.
[162] S. Rasmussen and D. Herfort. Mineralised Clinker and Cement. In: R.K. Dhir and TD
Dyer (eds.). Modern Concrete Materials: Binders, Addititions and Admixtures. Thomas
Telford, 1999, London. p. 59-67.
[163] F.L. Smidth Denmark A/S. Report on the Mini-RILEM method. Interal Report APr/EP,
1969.
[164] R. Kondo, M. Daimon, E. Sakai and H. Ushiyama. Influence of inorganic salts on the
hydration of tricalcium silicate. J. Appl. Chem. Biotechnol., 1977, 27, 191-197.
[165] X. Cong and R.J. Kirkpatrick.
29
Si and
17
O NMR investigation of the structure of some
crystalline calcium silicate hydrates. Advn. Cem. Based Mat., 1996, 3, 133-143.
[166] G.K. Sun, J.F. Young and R.J. Kirkpatrick. The role of Al in C-S-H: NMR, XRD, and
compositional results for precipitated samples. Cem. Concr. Res., 2006, 36, 18-29.
[167] M.D. Andersen, H.J. Jakobsen and J. Skibsted. Incorporation of aluminum in the calcium
silicate hydrate (CSH) of hydrated Portland cements: A high-field
27
Al and
29
Si MAS
NMR investigation. Inorg. Chem., 2003, 42, 2280-2287.
[168] F. Matsushita, Y. Aono and S. Shibata. Calcium silicate structure and carbonation
shrinkage of a tobermorite-based material. Cem. Concr. Res., 2004, 34, 1251-1257.
[169] J.J. Chen, J.J. Thomas, H.F.W. Taylor and H.M. Jennings. Solubility and structure of
calcium silicate hydrate. Cem. Concr. Res., 2004, 34, 1499-1519.
[170] I.G. Richardson. Tobermorite/jennite- and tobermorite/calcium hydroxide-based models
for the structure of C-S-H: applicability to hardened pastes of tricalcium silicate, |-
dicalcium silicate, Portland cement, and blends of Portland cement with blast-furnace
slag, metakaolin, or silica fume. Cem. Concr. Res., 2004, 34, 1733-1777.
References 139
[171] A. Popova, G. Geoffroy, M.F. Renou-Gonnord, P. Faucon and E. Gartner. Interactions
between polymeric dispersants and calcium silicate hydrates. J. Am. Ceram. Soc., 2000,
83, 2556-2560.
[172] H. Viallis, P. Faucon, J.C. Petit and A. Nonat. Interaction between salts (NaCl, CsCl
2
)
and calcium silicate hydrates (C-S-H). J. Phys. Chem. B, 1999, 103, 5212-5219.
[173] M.D. Andersen, H.J. Jakobsen and J. Skibsted. A new aluminium-hydrate species in
hydrated Portland cements characterized by
27
Al and
29
Si MAS NMR spectroscopy. Cem.
Concr. Res., 2006, 36, 3-17.
[174] M.D. Andersen. Structural investigations of Portland cement-based materials by solid-
state NMR spectroscopy. PhD Thesis 2004, University of Aarhus, Denmark.
[175] R. Taylor, I.G. Richardson and R. Brydson. Nature of C-S-H in 20 years old neat
ordinary Portland cement and 10% Portland cement - 90% ground granulated blast
furnace slag pastes. Adv. App. Ceram., 2007, 106, 294-301.
[176] A. Palomo, M.W. Grutzeck and M.T. Blanco. Alkali-activated fly ashes: a cement for the
future. Cem. Concr. Res., 1999, 29, 1323-1329.
[177] P. Duxson, A. Fernndez-Jimnez, J.L. Provis, G.C. Lukey, A. Palomo and J.S.J. van
Deventer. Geopolymer technology: the current state of art. J. Mater. Sci., 2007, 45, 2917-
2933.
[178] J. Davidovits. Geopolymer Chemistry and Applications. Institut Gopolymre, Saint-
Quentin, 2008.
[179] K. Komnitsas and D. Zaharaki. Geopolymer: A review and prospects for the minerals
industry. Minerals Engineering, 2007, 20, 1261-1277.
[180] P. Duxson, J.L. Provis and J.S.J. van Deventer. The role of inorganic polymer technology
in the development of 'green concrete'. Cem. Concr. Res., 2007, 37, 1590-1597.
[181] D.M.J. Sumajouw, D. Hardjito, S.E. Wallah and B.V. Rangan. Fly ash-based geopolymer
concrete: study of slender reinforced columns. J. Mater. Sci., 2007, 42, 3124-3130.
References 140
[182] M. Sofi, J.S.J. van Deventer, P.A. Demdis and G.C. Lukey. Engineering properties of
inorganic polymer concretes. Cem. Concr. Res., 2007, 37, 251-257.
[183] T.W. Cheng and J.P. Chiu. Fire-resistant geopolymer produced by granulated blast
furnace slag. Minerals Engineering, 2003, 3, 205-210.
[184] J.S.J. van Deventer, J.L. Provis, P. Duxson and G.C. Lukey. Reaction mechanisms in the
geopolymeric conversion of inorganic waste to useful products. J. Hazard. Mater., 2007,
139, 506-513.
[185] J. Davidovits. Geopolymers, Inorganic polymeric new materials. J. Therm. Anal., 1991,
37, 1633-1656.
[186] D. Khale and R. Chaudhary. Mechanism of geopolymerization and factors enfluencing its
development: a review. J. Mater. Sci., 2007, 42, 729-746.
[187] P.D. Silva, K. Sagoe-Crenstil and V. Sirivivatnanon. Kinetics of geopolymerization: Role
of Al
2
O
3
and SiO
2
. Cem. Concr. Res., 2007, 37, 512-518.
[188] T. Bakharev. Geopolymeric materials prepared using class F fly ash and elevated
temperature curing. Cem. Concr. Res., 2005, 35, 1224-1232.
[189] A.G. Gelevera and K. Munzer. Alkaline Portland and slag Portland cement. Proc. 1st
International Conference on Alkali Cement and Concrete 1, 1994, Kiev, Ukraine: p. 173-
179.
[190] Y.S.A. Fundi. Alkaline pozzolana Portland cement. Proc. 1st International Conference on
Alkali Cement and Concrete 1, 1994, Kiev, Ukraine: p. 181 - 192.
[191] A. Fernndez-Jimnez, A. Palomo, I. Sobrados and J. Sanz. The role played by the
reactive alumina content in the alkaline activation of fly ashes. Micropor. Mesopor.
Mater., 2006, 91, 111-119.
[192] M.R. Rowles, J.V. Hanna, K.J. Pike, M.E. Smith and H.H. O'Connor.
29
Si,
27
Al,
1
H and
23
Na MAS NMR Study of the bonding character in aluminosilicate inorganic polymers.
Appl. Magn. Reson., 2007, 32, 663-689.
References 141
[193] P. Duxson, J.L. Provis, G.C. Lukey, F. Separovic and J.S.J. van Deventer.
29
Si NMR study
of structural ordering in aluminosilicate geopolymer gels. Langmuir, 2005, 21, 3028-
3036.
[194] P.S. Singh, T. Bastow and M. Trigg. Structural studies of geopolymers by
29
Si and
27
Al
MAS NMR. J. Mater. Sci., 2005, 40, 3951-3961.
[195] J. Ding, Y. Fu and J.J. Beaudoin. Strtlingite formation in high alumina cement - silica
fume systems: Significance of sodium ions. Cem. Concr. Res., 1995, 25, 1311-1319.
[196] I.G. Lodeiro, A. Fernndez-Jimnez, A. Palomo and D.E. Macphee. Effect on fresh C-S-H
gels of the simultaneous addition of alkali and aluminium. Cem. Concr. Res., 2010, 40,
27-32.
[197] X. Pardal, I. Pochard and A. Nonat. Experimental study of SiAl substitution in calcium-
silicate-hydrate (C-S-H) prepared under equilibrium conditions. Cem. Concr. Res., 2009,
39, 637-643.
[198] R. Rinaldi, M. Sacerdoti and E. Passaglia. Strtlingite: crystal structure, chemistry, and a
reexamination of its polytype vertumnite. Eur. J. Mineral., 1990, 2, 841-849.
[199] S. Kwan, J. LaRosa and M.W. Grutzeck.
29
Si and
27
Al MASNMR study of strtlingite. J.
Am. Ceram. Soc., 1995, 78, 1921-1926.

References 142
Appendixes 143
Appendix 1

Sample preparations

Precursors
The white Portland clinker (HKL) was received from Aalborg Portland A/S, Denmark, and has
the following chemical composition: 70.30 wt. % CaO, 25.42 wt. % SiO
2
, 2.13 wt. % Al
2
O
3
,
0.37 wt. % Fe
2
O
3
, 0.63 wt. % MgO, 0.41 wt. % P
2
O
5
, 0.12 wt. % SO
3
, 0.04 wt. % F, 0.065 wt. %
K
2
O, 0.19 wt. % Na
2
O, 0.01 wt. % Cl, 0.094 wt. % TiO
2
, 0.0033 wt. % Cr
2
O
3
, 0.03 wt. % C and
a Blaine fineness of approx. 395 m
2
/kg.
Ca(OH)
2
:

Sigma-Aldrich Laborchemikalien GmbH, Seelze, Germany.
Gypsum: VWR International Ltd., Poole, England.
CaF
2
: Sigma-Aldrich Laborchemikalien GmbH, Seelze, Germany.
Na
2
SiF
6
: Merck KGaA, Darmstadt, Germany.
AlOOH: Boemitte, CONDEA Vista Company, Houston, Texas.
FeOOH: Riedel-De Han AG, Seelze Hannover.
SiO
2
: Silicagel, Fluka Chemie GmbH, Buchs, Switzerland.
Kaolinite: Fluka Chemie GmbH, Buchs, Switzerland.
NaAlO
2
: Strem Chemicals, New Buryport, USA.
NaSiO
3
9H
2
O: Fisher Chemicals, Fair Lawn, New Jersey.
10 M NaOH solution: Merck KGaA, Darmstadt, Germany.
KOH: Sigma-Aldrich Laborchemikalien GmbH, Seelze, Germany.
Sucrose: Merck KGaA, Darmstadt, Germany.


Fluoride-mineralized Portland clinker/cement
Clinker preparation
The fluoride-mineralized clinkers were synthesized using a white Portland clinker as the
main source. The desired bulk CaO, Al
2
O
3
, Fe
2
O
3
, SO
3
and F contents were achieved by adding
small quantities of Ca(OH)
2
, AlOOH, FeOOH, CaSO
4
and CaF
2
. Nodules with diameters of
about 1 cm were formed by hand from mixtures of typically 20 g powder and 7 ml water,
corresponding to a water/solid (w/s) ratio of 0.35. The wet nodules were dried at 200 C for one
day.
Appendixes 144
The clinkers were prepared by burning the dried nodules at 1450 ( 10) C for one hour in
a high-temperature furnace. Subsequently, the clinkers were cooled to 1250 1300 C and
immediately, quenched in water at ~25 C. Finally, the resulting clinkers were dried at 100 C for
one day and afterwards stored in closed containers.

Preparation of hydrated samples
The fluoride-mineralized clinkers were ground and sifted to a grain size below 40 m
prior to the hydration experiments. The hydrated samples were prepared from a mixture of water
(w/c = 0.5), gypsum (SO
3
/Al
2
O
3
= 1.3) and typically 10 g clinker. The paste was mixed by hand
for 15 minutes and subsequently stored over water (RH = 100 %) in a sealed desiccator at room
temperature.
The hydration was stopped after 6h, 12h, 1, 2, 7, 28, 90 and 200 days. Initially, a small
piece (~1.0 g) was knocked off the paste cylinder and ground to a fine powder. Subsequently, the
powder was suspended in acetone for 15 minutes under slow rotation using a magnetic stirrer.
Finally, the powder was separated from the acetone and dried at room temperature in a
desiccator for 24 hours. The samples were stored in small sealed specimen tubes to avoid
reactions with moisture and CO
2
.

Selective dissolution
A solution containing 15 g KOH, 15 g sucrose and 150 ml water was heated to 90 C. A
quantity of 5 g clinker was added to this solution and stirred for 1 minute. The residue was then
washed in 50 ml water and 100 ml methanol and dried at 60 C.

Preparation of CSH samples
The CSH samples were prepared from typically 2 g powder mixture of Ca(OH)
2
, SiO
2
,
AlOOH and Na
2
SiF
6
. The solids, with the desired composition, were mixed with three-times
distilled water using a water/solid ratio of 45. The mixture was stored at room temperature for
three weeks in closed glass tubes with continuous stirring using a magnetic stirrer. The reaction
was stopped by filtering the suspensions and the precipitates were washed with three-time
distilled water. The resulting products were dried in a desiccator over silica gel at room
temperature.

Appendixes 145
Alkali-activated metakaolin
Metakaolin was prepared by heat treatment of kaolinite at 600
o
C for 4 hours. The powder
was cooled to room temperature in a desiccator before adding the alkali-activating solution. The
desired molar Si/Al and Na/Al ratios were obtained by adjusting the quantities of metakaolin,
silica and the amount of 10 M NaOH solution. Silica was mixed with the 10 M NaOH solution
prior to the addition of metakaolin. The resulting mixtures were cured at 80
o
C in closed glass
tubes and the reactions were stopped after one week by suspending the mixtures in acetone for
ca. 15 minutes. Finally, the samples were dried in a desiccator, ground, and stored in a closed
container to avoid reactions with moisture and CO
2
.


Synthetic strtlingite
Strtlingite were prepared from a mixture of NaSiO
3
9H
2
O, Ca(OH)
2
and NaAlO
2
in
accordance with its stoichiometry of 2CaOAl
2
O
3
SiO
2
8H
2
O. The powder was homogenously
mixed with water using a w/s ratio of 10. Subsequently, the mixture was cured in an ultrasound
bath for 53 days at room temperature. Finally, the powder was filtrated from the solution and
dried at 100 C for one day.

Appendixes 146
Appendix 2

NMR measurements and other analytical techniques
The Instrument Centre for Solid-State NMR Spectroscopy has four Varian NMR
spectrometers with different magnetic fields: Unity INOVA-200 (4.7 T), Unity INOVA-300 (7.1
T), Unity INOVA-400 (9.4 T) and Direct Drive VNMRS-600 (14.1 T).

29
Si MAS NMR experiments
The
29
Si MAS NMR spectra (7.1 T) were recorded using a home-built CP/MAS probe for
7 mm o.d. rotors, a pulse width of 3 s for an rf-field strength of v
Si
=
Si
B
1
/2t = 58 kHz, a
spinning speed of v
R
= 7.0 kHz, a 30-s relaxation delay and typically 2048 scans.
The
29
Si{
19
F} CP-REDOR NMR experiments (7.1 T) for the fluoride-mineralized
Portland clinkers used a home-built CP/MAS probe for 5 mm o.d. rotors, v
R
= 10.0 kHz, a 8-s
relaxation delay and 10560 scans. The HH-matching was obtained for v
Si
= 38 kHz, v
F
= 47 kHz
and the CP part used a decreasing linear ramped amplitude (RAMP) on the
19
F channel with an
amplitude variation of 10 kHz and a CP contact time of 3.0 ms. The rf-field strengths (v
rf
=
B
rf
/2t) applied in the REDOR sequence are v
Si
= 38 kHz and v
F
= 67 kHz, corresponding to t-
pulses of 13.2 and 7.5 s, respectively. The
29
Si{
19
F} CP/MAS experiments for this series of
samples were performed with the same conditions but used a spinning speed of 5.0 kHz, v
Si
=
47.5 kHz and v
F
= 42.1 kHz.
The
29
Si{
19
F} CP/MAS NMR spectra (7.1 T) of the synthetic CSH samples were
recorded using a home-built CP/MAS probe for 7 mm o.d. rotors, v
R
= 10.0 kHz, a 10-s
relaxation delay and 16,384 22,400 scans. The HH-matching was obtained for v
Si
= 41.0 kHz,
v
F
= 31.2 kHz and the CP part used a RAMP pulse of 10 kHz on the
19
F channel for a CP contact
time of 5.0 ms.
The
29
Si{
27
Al} REAPDOR NMR experiments (14.1 T) used a triple-resonance MAS
probe for 5 mm o.d. rotors from DOTY Scientific Inc. The spectra were acquired using rf-field
strengths of v
Si
= 49 kHz (t/2-pulse = 5.1 s and t-pulse = 10.2 s) and v
Al
= 50 kHz, a
spinning speed of v
R
= 10.0 kHz, a 30-s relaxation delay and typically 960 scans for the
individual spectra without (S
0
) and with (S)
27
Al irradiation.

Appendixes 147
27
Al MAS NMR experiment
27
Al MAS NMR spectra (14.1 T) were recorded using a home-built CP/MAS probe for 4
mm o.d. rotors, a pulse width of 0.5 s for an rf-field strength of v
Al
= 60 kHz to ensure
quantitative reliability of the intensities observed for the
27
Al central transitions for sites
experiencing different quadrupole couplings.

The
27
Al NMR experiments typically employed
1
H
decoupling with v
H
= 50 kHz, a spinning speed of 13.0 kHz, a 2-s relaxation delay and ~3,200
scans. Furthermore, the
27
Al MAS NMR spectrum of the probe itself with an empty spinning
PSZ rotor showed a broadened resonance, which was subtracted from the
27
Al MAS NMR
spectra of the samples prior to the evaluation of spectra.
The triple-quantum
27
Al MQMAS experiment (9.4 T) was performed using the three-
pulse z-filter sequence with an rf-field strength of B
1
/2t = 60 kHz for the first and second
pulses and B
1
/2t = 25 kHz for third selective 90 pulse. Furthermore,
1
H decoupling (B
2
/2t =
40 kHz) was employed in the evolution and detection periods.
The
27
Al{
19
F} CP/MAS NMR experiment (7.1 T) was performed using a home-built
CP/MAS probe for 7 mm o.d. rotors. The Hartmann-Hahn (HH) match was obtained for v
Al
=
12.5 kHz and v
F
= 35.8 kHz using a CP contact time of 1.5 ms, v
R
= 5.0 kHz, a relaxation delay
of 4 s and 72,384 scans. The CP experiment was further improved using a decreasing linear
ramped amplitude on the
19
F channel with an amplitude variation of 15 kHz on the
19
F channel
during the CP contact time.

19
F MAS NMR MAS experiments
The
19
F MAS NMR spectra (7.1 T) were recorded using a home-built CP/MAS probe for
7 mm o.d. rotors, a pulse width of 5 s for an rf-field strength of v
F
= 50 kHz, v
R
= 10.0 kHz, a
10-s relaxation delay and typically 512 scans for the anhydrous samples, 2048 8192 for
hydrated Portland cements and 128 for the synthetic CSH.
The
19
F MAS NMR spectra for the synthetic CSH samples recorded at 14.1 T used a
home-built CP/MAS probe for 4 mm o.d. rotors, a pulse width of 5.6 s for an rf-field strength
of 48 kHz, a v
R
= 13.0 kHz, a 15-s relaxation delay and typically 512 scans.

Appendixes 148
43
Ca MAS NMR experiments
The
43
Ca MAS NMR experiments were performed at 14.1 T used two different probes: a
home-built CP/MAS probe for 7 mm o.d. rotors and a Chemmagnetic probe for 7.5 o.d. rotor.
The Hartmann-Hahn (HH) match for the
43
Ca{
19
F} CP/MAS NMR spectrum was obtained for

Al
B
1
/2t = 13 kHz and
F
B
2
/2t = 36 kHz using a CP contact time of 1.5 ms, v
R
= 5.0 kHz, a
relaxation delay of 4 s and 72,384 scans.
The
19
F,
27
Al,
29
Si and
43
Ca chemical shifts were referenced to external samples of neat
tetramethylsilane (TMS) using a secondary reference sample of |-Ca
2
SiO
4
(71.33 ppm), a 1.0
M AlCl
3
.
6H
2
O solution, CFCl
3
using a secondary reference sample of Na
2
SiF
6
(-149.3 ppm) and
a 1.0 M CaCl
2
solution, respectively.

Other analyses including measurements of the free lime content, the bulk fluoride content
and the chemical compositions by XRF and XRD Rietveld analyses have been conducted at the
chemical laboratory at Aalborg Portland A/S.

Free lime content measurement
The free lime content was measured by dissolving the clinker in ethylene glycol.
Subsequently, the solution was titrated with hydrochloric acid.

Bulk fluoride content measurement
The clinkers were dissolved in a solution of water, HCl and alum KAl(SO
4
)
2
. The
fluoride concentration in the resulting mixture was measured by an ion selective electrode. The
solvent were prepared by dissolving 16 g alum in 800 ml water and 800 ml conc. HCl and
finally, diluted with water to 2.0 litres.

Appendixes 149
Procedure for fitting REDOR data in Mathematica
Data
data = {{x1,y1},{x2,y2}...,{xn,yn}}
g1 = 2.518148*10^8
g2 = 5.3190*10^7
u = 4Pi*10^(-7)
h = 1.0546*10^(-34)
spin = 10000
k =

Fitting
<< Statistics`NonlinearFit`
sred[n_,dtr_,k_] := 1. - BesselJ[0,Sqrt[2.]*n*dtr]^2. + Sum[2.BesselJ[i,Sqrt[2.]*n*dtr]^2./(16. i^2. - 1.0),
{i, 1, k}]
besselfit[k_,init_] := NonlinearFit[data, sred[n,dtr,k], n, {dtr, init}, AccuracyGoal\[Rule]Automatic,
PrecisionGoal\[Rule]Automatic, ShowProgress\[Rule]True]
besselfit[k,0.12]

Plotting
f[n_]:= 1. - BesselJ[0,Sqrt[2.]*n*dtr]^2. + Sum[2. BesselJ[i,
n*Sqrt[2.0]*dtr]^2. /((16. i^2. - 1.0)),{i,1,k}])
pnkt=ListPlot[data,PlotStyle\[Rule]{RGBColor[1,0,0],Thickness[1]}]
graf1=Plot[f[n],{n,0,80},PlotStyle\[Rule]{RGBColor[0,0,1]},Frame\[Rule]True,FrameLabel\
[Rule]{"nTr",\[CapitalDelta]S/Subscript[S,0]},RotateLabel\[Rule]False,PlotRange\[Rule]{0,1.1},
TextStyle\[Rule]{FontSize\[Rule]20}]
Show[graf1,pnkt]
Export["image.bmp",Show[graf1,pnkt],ImageSize\[Rule]1000]

Calculating the distance
dtr=Dipol=dtr/(1/spin)
r=(g1*g2*h*u/(2Pi*4Pi*Dipol))^(1/3)
Appendixes 150

Potrebbero piacerti anche