205
Cationic and Anionic Ordering in Tetrahedral and Octahedral Sheets of Synthetic Al-rich Phlogopite Investigated by Solid-State NMR Spectroscopy and Monte-Carlo Simulations Dissertation zur Erlangung des Grades der Doktorwürde der Naturwissenschaften Dr. rer. nat. der Fakultät für Geowissenschaften der Ruhr-Universität Bochum vorgelegt von Dipl.-Min. Ramona Langner aus Rochlitz im Juli 2010

Cationic and Anionic Ordering in Tetrahedral and ... · Cationic and Anionic Ordering in Tetrahedral and Octahedral Sheets of Synthetic Al-rich Phlogopite Investigated by Solid-State

  • Upload
    doduong

  • View
    233

  • Download
    0

Embed Size (px)

Citation preview

Cationic and Anionic Ordering in Tetrahedral

and Octahedral Sheets of Synthetic Al-rich

Phlogopite Investigated by Solid-State NMR

Spectroscopy and Monte-Carlo Simulations

Dissertation

zur Erlangung des Grades der Doktorwürde

der Naturwissenschaften

Dr. rer. nat.

der Fakultät für Geowissenschaften

der Ruhr-Universität Bochum

vorgelegt von

Dipl.-Min. Ramona Langner

aus Rochlitz

im Juli 2010

Erster Gutachter: Priv.-Doz. Dr. Michael Fechtelkord

Zweiter Gutachter: Dr. Alberto García Arribas

Fachfremder Gutachter: Prof. Dr. Harald Zepp

Datum der Abgabe: 14.07.2010

Datum der Disputation: 26.10.2010

Abstract

II

Abstract

The aim of this study was to investigate the relationship between cation and

anion ordering in the tetrahedral and octahedral sheets of the mica phlogopite.

Phlogopite samples have been synthesised at T = 600 °C and p = 2 kbar with a

run duration of one week. A wide range of nominal compositions

K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y has been covered, reaching from F-free to

water-free compositions (0.0 ≤ y ≤ 2.0) and from Mg-phlogopites to very Al-rich

samples (0.0 ≤ x ≤ 1.6). The obtained samples have been investigates using solid-

state NMR spectroscopy, powder X-ray diffraction and scanning electron microscopy.

At the same time, Monte-Carlo simulations have been performed by Dr. Alberto

García Arribas, Institut de Ciència de Materials de Barcelona, CSIC, Bellaterra,

Spain, and Dr. Javier López-Solano, Universidad del Pais Vasco, Bilbao, Spain, for

F-free compositions. These simulations were based on the so-called ‘J-formalism’

describing ordering in terms of exchange reactions concerning neighbouring sites.

The ordering patterns found in the atomic configurations obtained from the

simulations were then compared to the experimental results.

For all compositions, the Al-content of the tetrahedral sheets estimated from 29Si

MAS NMR spectra has been found to be lower than the Al-content of the initial oxide

mixtures. The highest amount of Al incorporated into the structure was observed for

hydroxyl-phlogopites (y = 2.0). For nominal compositions of xnom = 1.0 and 1.2 the

estimated Al-content xest was 0.83. At even higher initial Al-contents (xnom = 1.6) the

amount of Al incorporated decreased again.

Phlogopites containing fluorine showed a reduced ability to incorporate Al into

their crystal structure. As soon as F was added to the initial oxide mixture the

estimated Al-content of the phlogopites decreased considerably compared to F-free

compositions. However, the exact amount of F was not significant as for a given

nominal Al-content the amount of Al incorporated into the structure has been roughly

the same for all F-contents. Only for water-free systems a sharp decrease in xest has

been observed again. It can be concluded that the mere presence of F in the mixture

has a much stronger influence on the phlogopites’ ability to incorporate Al than the

exact ratio of OH/F.

Abstract

III

Excess Al led to the formation of impurity phases. Except for samples of very low

initial Al-contents, aluminium oxide (Al2O3) has been observed for all compositions.

Another impurity phase was potassium aluminium hexafluoride (K3AlF6*0.5H2O)

which was not only formed at high F- and Al-contents, but also in samples containing

only small amounts of F if the Al-content was high enough. The formation of kalsilite

has not been observed in the samples studied here.

The ordering of tetrahedral cations is dominated by next-nearest-neighbour

interactions. The corresponding interaction parameter 1J has been found to be highly

positive which means occupation of directly neighboured tetrahedra by two Al-atoms

is avoided. At a maximum Al-content of x = 1.0, long-range ordering occurs with Al

and Si occupying tetrahedra alternately. At lower amounts of Al a separation into two

areas of different compositions has been observed in the configurations of lowest

energy obtained from Monte Carlo simulations. On the one hand the ordered

structure is preserved in clusters of composition Si/[4]Al = 1:1. On the other hand the

lower Al-content is compensated by the formation of clusters showing a composition

similar to phlogopite in the narrower sense, i.e. without additional tetrahedral Al

(Si/[4]Al = 3:1). The Al-poor clusters are characterised by short-range ordering

controlled by the avoidance of [4]Al-O-[4]Al linkages.

19F and 1H MAS NMR spectroscopy has been applied to investigate the ordering

of Mg/Al and OH/F in the octahedral sheets. A strong preference of F for a co-

ordination by 3 Mg and of OH for 2MgAl has been observed, respectively. For

hydroxyl-groups this preference decreases with increasing nominal Al-content of the

samples. In contrast, the amount of F being co-ordinated by Mg only increases for

higher initial Al-contents.

Octahedral cations have been found to be completely ordered for x = 1.0. [6]Al is

always co-ordinated by six Mg-atoms, therefore avoiding a direct neighbourhood of

two Al-atoms in adjacent octahedra, similar to [4]Al-O-[4]Al avoidance in the tetrahedral

sheet. Again, clustering is observed for lower Al-contents. Ordered clusters of a

composition with Mg/[6]Al = 2:1 are separated by areas containing only Mg.

The relationship between both ordering patterns has also been investigated.

{1H} → 29Si HETCOR NMR experiments revealed a close neighbourhood of Al-rich

tetrahedral and octahedral clusters, and this has been confirmed by the simulation

results. Two Al-atoms occupying directly neighboured sites has been observed to be

Abstract

IV

favourable if two different types of polyhedra are involved. Moreover, {1H/19F} → 29Si

CPMAS NMR experiments showed a clustering of OH and F in the octahedral sheet.

F-rich octahedral environments are more likely to be found near Al-poor tetrahedral

areas. It can be concluded that the clustering involves all sheets of a single layer

package, leading to a separation of clusters of the two end-member compositions

K Mg (AlSi3O10) F2 (fluoro-phlogopite) and K (Mg2Al) (Al2Si2O10) (OH)2 (‘eastonite’),

respectively.

Powder X-ray diffraction patterns showed that the exchange of Mg/Si by [6]Al/[4]Al

not only influences the local atomic arrangement, but also affects the crystallisation of

polytypes. Structural changes resulting from the different cationic radii of Mg2+ and

Al3+ lead to the formation phlogopite-2M1 next to the 1M-polytype which is more

common in natural samples. These two polytypes are intergrown leading to stacking

faults and a high degree of disorder in the whole structure.

Kurzfassung

V

Kurzfassung

Ziel der vorliegenden Arbeit war die Aufklärung des Zusammenhangs zwischen

Kationen- und Anionenordnung in den Tetraeder- und Oktaederschichten des

Glimmers Phlogopit.

Dazu wurden Phlogopit-Proben bei einer Temperatur von 600 °C und einem

Druck von 2 kbar sowie einer Synthesedauer von einer Woche hergestellt. Ein großer

Bereich an nominellen Zusammensetzungen K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y

wurde abgedeckt, der von fluor- zu wasserfreien Zusammensetzungen reichte (0.0 ≤

y ≤ 2.0) und sowohl Al-freie als auch Al-reiche Proben umfasste (0.0 ≤ x ≤ 1.6). Die

auf diese Weise erhaltenen Proben wurden mittels Festkörper-NMR-Spektroskopie,

Röntgenpulverdiffraktion und Rasterelektronenmikroskopie untersucht.

Zugleich führten Dr. Alberto García Arribas, Institut de Ciència de Materials de

Barcelona, CSIC, Bellaterra, Spanien, und Dr. Javier López-Solano, Universidad del

Pais Vasco, Bilbao, Spanien, Monte-Carlo-Simulationsrechnungen für F-freie

Zusammensetzungen durch, basierend auf dem sogenannten „J-Formalismus“.

Dieser beschreibt Ordnung auf benachbarten Kationenplätzen durch

Austauschreaktionen. Die Ordnungsmuster in den durch die Simulationen erhaltenen

atomaren Konfigurationen konnten dann mit den experimentellen Ergebnissen

verglichen werden.

In allen Proben war der mithilfe von 29Si MAS NMR-Spektren bestimmte Al-

Gehalt der Phlogopite geringer als der der Ausgangsmischung. Reine Hydroxyl-

Phlogopite zeigten den höchsten Al-Gehalt: Für eine nominelle Zusammensetzung

von xnom = 1.0 und 1.2 wurde ein geschätzer Wert xest = 0.82 gefunden. Ein

Sättigungseffekt wurde insofern beobachtet, als dass bei noch höherem nominellen

Al-Gehalt die Menge an eingebautem Al wieder abnahm.

Im Vergleich dazu zeigten F-haltige Phlogopite ein geringeres Vermögen, Al in

ihre Struktur aufzunehmen. Für Proben unterschiedlichen F-Gehalts war die Menge

an aufgenommenem Al für gleiche nominelle Zusammensetzung jedoch ähnlich.

Lediglich bei wasserfreien Proben zeigte sich eine deutliche Abnahme des Al-

Gehalts bei gleichem nominellem Al-Gehalt. Daraus folgt, dass die Präsenz von F in

der Ausgangsmischung einen deutlichen Effekt auf das Vermögen der Phlogopite Al

aufzunehmen hat, während der genaue F-Gehalt keine so wichtige Rolle spielt.

Kurzfassung

VI

Das überschüssige Aluminium führte zur Bildung von Nebenphasen. Mit

Ausnahme von Proben mit sehr niedrigem nominellen Al-Gehalt wurde für alle

Zusammensetzungen Aluminiumoxid (Al2O3) gefunden. Eine weitere Nebenphase

war Kaliumaluminiumhexafluorid (K3AlF6*0.5H2O), welches nicht nur bei hohen F-

Gehalten auftrat, sondern sich auch bei zwar F-armen aber sehr Al-reichen

Zusammensetzungen bildete.

Die Kationenordnung der Tetraederschicht wird von der Wechselwirkung

zwischen nächsten Nachbarn dominiert. Der hohe positive Wert des 1J -Parameters,

der die zugehörige Wechselwirkung beschreibt, führt zu einer Vermeidung von zwei

Al-Atomen auf benachbarten Tetraederplätzen. Beim maximalen Al-Gehalt von

x = 1.0 herrscht langreichweitige Ordnung vor, bei der Si und Al abwechselnd

Tetraederpositionen besetzen. Bei niedrigeren Al-Gehalten wurde eine Zweiteilung

der Struktur beobachtet: Einerseits bleibt die geordnete Struktur in Clustern mit einer

Zusammensetzung von Si/[4]Al = 1:1 erhalten, andererseits wird der geringere Gehalt

an Aluminium durch die Ausbildung von Clustern aus Phlogopit im engeren Sinne

ausgeglichen, in denen das Verhältnis von Si/[4]Al 1:1 beträgt. Letztere weisen nur

noch Ordnung von kurzer Reichweite auf, bedingt durch die Vermeidung von [4]Al-O-[4]Al-Bindungen im Sinne der Loewenstein’schen Regel.

Die Ordnung von Mg/Al und OH/F in der Oktaederschicht wurde mithilfe von 19F

und 1H MAS NMR-Spektroskopie untersucht. Es zeigte sich, dass F eine

Koordination durch 3 Mg bevorzugt, während OH eine Mg2Al-Umgebung vorzieht. Mit

zunehmender Menge an eingebautem Al nähert sich der Al-Gehalt der OH-

Umgebungen dem der gesamten Oktaederschicht an. Im Gegensatz dazu steigt die

Neigung des Fluors, Plätze mit reiner Mg-Umgebung aufzusuchen, bei höheren Al-

Gehalten.

Auch die Kationen der Oktaederschicht weisen für eine Zusammensetzung von

xnom = 1.0 eine Ordnung von langer Reichweite auf. Dabei wird Al stets von sechs

Mg-Atomen auf den benachbarten Plätzen umgeben, um zu vermeiden das Al-Atome

als nächste Nachbarn auftreten. Ähnlich den Verhältnissen in der Tetraederschicht

wurde auch hier die Ausbildung von Clustern beobachtet. Dabei treten Cluster mit

einer Zusammensetzung von Mg/[6]Al = 2:1 neben Al-freien Clustern auf.

Der Zusammenhang zwischen beiden Ordnungsmustern wurde ebenfalls

untersucht. Mithilfe von {1H} → 29Si HETCOR NMR-Experimenten konnte gezeigt

Kurzfassung

VII

werden, dass solche Cluster der Tetraederschicht, die einen hohen Al-Gehalt

aufweisen, sich in direkter Nachbarschaft von gleichfalls Al-reichen Clustern der

Oktaederschicht befinden. Dies wurde auch durch Monte-Carlo-Simulationen

bestätigt. Das bedeutet, dass die Besetzung zweier benachbarter Plätze durch zwei

Al-Atome doch energetisch günstig sein kann, wenn es sich dabei um zwei

verschiedene Typen von Koordinationspolyedern handelt. Darüber hinaus

bestätigten {1H/19F} → 29Si CPMAS-Experimente, dass auch OH und F jeweils in

Clustern angeordnet sind. F-reiche Cluster befinden sich dabei in nächster Nähe zu

Al-armen Bereichen in der Tetraederschicht. Daraus folgt, dass die Clusterbildung

alle Schichten eines einzelnen Schichtpaketes umfasst und dabei eine Trennung in

Cluster der zwei Endglied-Zusammensetzungen K Mg (AlSi3O10) F2 (Fluoro-

Phlogopit) und K (Mg2Al) (Al2Si2O10) (OH)2 (‚Eastonite’) stattfindet.

Röntgenpulverdiffraktogramme zeigten weiterhin, dass der Austausch von Mg/Si

durch [6]Al/[4]Al nicht nur die atomare Umgebung beeinflusst, sondern sich auch auf

die Kristallisation verschiedener Polytype auswirkt. Bedingt durch unterschiedliche

Kationenradien von Al3+ und Mg2+ führen strukturelle Veränderungen zur Bildung von

Phlogopit-2M1 neben dem in natürlichen Proben bedeutenderen 1M-Polytyp. Beide

Polytype sind miteinander verwachsen, was zu Stapelfehlern und Fehlordnung in der

gesamten Kristallstruktur führt.

IX

Table of Contents

ABSTRACT ................................................................................................................ II

KURZFASSUNG ........................................................................................................ V

1. INTRODUCTION .................................................................................................... 1

2. THEORY ................................................................................................................. 7

2.1. Phlogopite structure and mineralogy ........................................................................................... 7 2.1.1 General chemical composition of micas .................................................................................... 7 2.1.2. The phlogopite structure ........................................................................................................... 8 2.1.3. Polytypism ............................................................................................................................... 10 2.1.4. Ordering of tetrahedral cations ............................................................................................... 14 2.1.5. Ordering of octahedral cations ................................................................................................ 15 2.1.6. Exchange of OH by F in the octahedral sheet ........................................................................ 17 2.1.7. Phlogopite mineralogy ............................................................................................................. 18

2.2 Solid-state NMR spectroscopy .................................................................................................... 20 2.2.1. Interactions influencing NMR lineshapes ................................................................................ 21

2.2.1.1. Zeeman interaction .......................................................................................................... 21 2.2.1.2. Chemical shift interaction ................................................................................................. 22 2.2.1.3. Dipolar interaction ............................................................................................................ 23 2.2.1.4. Quadrupolar interaction ................................................................................................... 26

2.2.2. Experimental techniques ......................................................................................................... 28 2.2.2.1. Magic angle spinning NMR spectroscopy ....................................................................... 28 2.2.2.2. Cross-polarisation magic angle spinning NMR spectroscopy ......................................... 29 2.2.2.3. 2D hetero-nuclear correlation CPMAS NMR experiments .............................................. 34 2.2.2.4. Multiple quantum MAS NMR spectroscopy ..................................................................... 35

2.3 J-formalism and Monte-Carlo simulations .................................................................................. 40

3. EXPERIMENTAL AND ANALYTICAL METHODS .............................................. 47

3.1. General approach ......................................................................................................................... 47

3.2 Sample preparation ....................................................................................................................... 50 3.2.1. Preparation of gels .................................................................................................................. 50 3.2.2. Hydrothermal synthesis ........................................................................................................... 50

3.3. NMR spectroscopic experiments ................................................................................................ 52 3.3.1. 1H MAS NMR experiments ...................................................................................................... 52 3.3.2. 29Si MAS NMR experiments .................................................................................................... 52 3.3.3. 27Al MAS NMR and 27Al 3QMAS NMR experiments ............................................................... 52 3.3.4. 19F MAS NMR experiments ..................................................................................................... 53 3.3.5. 17O MAS and 17O MQMAS NMR experiments ........................................................................ 54 3.3.6. {1H} → 29Si CPMAS/HETCOR experiments ............................................................................ 55 3.3.7. {19F} → 29Si CPMAS/HETCOR experiments ........................................................................... 55

3.4. X-ray diffraction experiments ...................................................................................................... 56

3.5. Scanning electron microscopy ................................................................................................... 57

X

4. RESULTS AND DISCUSSION ............................................................................. 59

4.1. General description of samples .................................................................................................. 59

4.2. Ordering of cations in the tetrahedral sheets of phlogopite .................................................... 64 4.2.1. Samples of intermediate to low F-contents (1.0 ≤ y ≤1.8) ...................................................... 64 4.2.2. Hydroxyl-phlogopites and Al- and OH-rich phlogopites (0.8 ≤ x ≤ 1.6; 1.6 ≤ y ≤2.0) .............. 68 4.2.3. Samples of high F-contents (y < 1.0) ...................................................................................... 72 4.2.4. J-formalism and Monte-Carlo simulations ............................................................................... 74

4.3. Ordering of cations and anions in the octahedral sheets of phlogopite ................................ 82 4.3.1. Samples of intermediate to low F-contents (1.0 ≤ y ≤ 1.8) ..................................................... 82

4.3.1.1. 1H MAS NMR spectroscopy ............................................................................................. 82 4.3.1.2. 19F MAS NMR .................................................................................................................. 86

4.3.2. Hydroxyl-phlogopites and Al- and OH-rich phlogopites .......................................................... 89 (0.8 ≤ x ≤ 1.6; 1.6 ≤ y ≤2.0) ............................................................................................................... 89

4.3.2.1. 1H MAS NMR spectroscopy ............................................................................................. 89 4.3.2.2. 19F MAS NMR spectroscopy ............................................................................................ 93

4.3.3. Samples of high F-contents (y < 1.0) ...................................................................................... 97 4.3.3.1. 1H MAS NMR spectroscopy ............................................................................................. 97 4.3.3.2. 19F MAS NMR spectroscopy ............................................................................................ 97

4.3.4. J-formalism and Monte-Carlo simulations ............................................................................. 101

4.4. Relationship between the ordering of ions in the tetrahedral and in the octahedral sheets of phlogopite .......................................................................................................................................... 106

4.4.1. Hydroxyl-phlogopites (y = 2.0) .............................................................................................. 106 4.4.1.1. 2D {1H} → 29Si HETCOR CPMAS NMR spectroscopy .................................................. 106 4.4.1.2. J-formalism and Monte-Carlo simulations ..................................................................... 108

4.4.2. F-containing phlogopites ....................................................................................................... 114

4.5. 27Al MAS and 27Al 3Q-MAS NMR spectroscopy ....................................................................... 122

4.6. 17O MAS and 17O 3Q-MAS NMR spectroscopy ........................................................................ 134

4.7. Analysis of X-ray diffraction powder patterns ........................................................................ 139

5. CONCLUSIONS AND OUTLOOK ..................................................................... 147

A. APPENDIX ......................................................................................................... 150

A.1. List of abbreviations .................................................................................................................. 150

A.2. NMR spectroscopic results ...................................................................................................... 153

B. REFERENCES .................................................................................................. 171

LIST OF TABLES .................................................................................................. 181

LIST OF FIGURES ................................................................................................. 183

DANKSAGUNG ..................................................................................................... 189

LEBENSLAUF ....................................................................................................... 191

ERKLÄRUNG ......................................................................................................... 193

1. Introduction

1

1. Introduction

Micas are a class of widely distributed minerals, formed in virtually all types of

rocks under varying conditions. They are present in sediments and sedimentary

rocks on Earth’s surface and remain stable through all fields of metamorphic rocks

down to the lower crust. In rocks like kimberlites which are thought to originate from

the mantle, micas occur next to high-pressure minerals like diamond. These minerals

also crystallise in many types of plutonic and volcanic igneous rocks. Especially in

granitic pegmatites huge crystals of several meters in diameter may occur

(Rickwood, 1981).

Accordingly, a large number of publications dedicated to micas have been

published so far, and micas in general have been reviewed extensively in 1984

(Reviews in Mineralogy 13, Bailey, Ed.) and a second time in 2002 (Reviews in

Mineralogy 46, Mottana et al., Eds.). Nevertheless, there is only little knowledge

about their structural details and their chemical composition and stability. This is

partly due to the complex crystal chemistry of this mineral class: A large variety of

ions may be incorporated into their crystal structure and excessive exchange

reactions may take place, as they form under a wide range of pressure and

temperature conditions. On the other hand, their structural disorder and the layered

texture make it difficult to obtain samples suitable for detailed structural analysis.

However, additional knowledge about micas may be useful to obtain information

on the formation conditions of metamorphic and igneous rocks, on the release of

water and the resulting production of melt in the lower crust and upper mantle, the

alteration of sediments and the formations of soils. Due to its extraordinary capability

to incorporate larger amounts of F than most other minerals, this is especially true for

phlogopite, the Mg-end-member of the biotite solid-solution series (K Mg3 (AlSi3O10)

(OH,F)2). Within the class of mica minerals this is only exceeded by the Li-mica

lepidolite (Foster, 1960). Besides, not many other minerals have been found to obtain

considerable amounts of F.

F is often present in silicic magmas only in minor amounts, but it may be strongly

enriched in the melt during ongoing crystallisation because of its incompatible

character. As a result, F-rich minerals like phlogopite form in late-stage magmatic

rocks like pegmatites (e.g., Christiansen et al., 1983; London, 1987). For certain A-

1. Introduction

2

type granites, F-contents up to 1.8 wt% have been found (Whalen et al., 1987), and

even larger amounts of 3.2 wt% F have been reported for topaz rhyolites by

Pichavant and Manning (1984). These amounts may have a strong influence on the

physical and chemical properties of magma with effects similar to those of water

solved in the melt. F lowers the crystallisation temperature of a melt (Manning, 1981;

Webster et al., 1987; Weidner and Martin, 1987), it decreases melt density (Dingwell

et al., 1993; Knoche et al., 1995) and melt viscosity (Dingwell et al., 1985; Baker and

Vaillancourt, 1995; Giordano et al., 2004), and increases element diffusivity in the

melt (Baker and Bossànyi, 1994). However, there is an important difference in the

behaviour of F and H2O: The water solubility decreases upon ascent of the magma,

leading to a higher viscosity and higher solidus temperatures, and thus a more

explosive nature of eruptions. In contrast, the fluorine solubility may still achieve

several wt% of fluorine even at low pressures, inhibiting degassing upon extrusion,

corresponding to a completely different behaviour of the melt (Carroll and Webster,

1994)

Therefore, it is essential to gain a deeper understanding of the stability of such F-

rich minerals and the processes controlling a partitioning of F between mineral and

co-existing melt. This includes studies of phase equilibria, partitioning coefficients

and thermal stability of micas. However, it is also necessary to obtain further

information on the local F-environment in the melt as well as in the F-containing

crystal structures. In contrast to standard techniques like X-ray and neutron

diffraction, spectroscopic methods are ideal tools to obtain information on the local

environment of single atoms in the structure.

In micas, F-incorporation is strongly related to the Al-content of the minerals: The

higher the Al-content, the less the mica’s ability to replace OH by F. Phlogopite in the

narrower sense does not contain any octahedral Al, but natural phlogopite crystals

always contain additional Al in the octahedral as well as the tetrahedral sheet. The

composition then ranges towards the hypothetical end-member ‘eastonite’ (K (Mg2Al)

(Al2Si2O10) (OH,F)2). Therefore, the investigation of both elements in the phlogopite

structure cannot be undertaken separately.

The relationship between Al-content and F-incorporation has first been described

in detail by Robert and Kodama (1988) in their IR-spectroscopic study of trioctahedral

micas. These authors observed a weaking of the interaction of OH-groups with apical

1. Introduction

3

oxygen atoms of (Si,Al)O4-tetrahedra at lower overall Al-contents. This effect led to

an increase in K+-H+-repulsion and allowed for an easy substitution of OH by F. In

contrast, higher Al-contents strengthened the OH…O interaction and lowered the

K+-H+-repulsion, and F was only incorporated into the structure in limited amounts.

Further studies showed that the F-Al avoidance was not only present on a

macroscopic scale but also on the atomic level. Huve et al. (1992a,b) studied the F-

environment in several natural and synthetic layer silicates with 19F MAS NMR

spectroscopy and observed a relationship between the position of the signal resulting

from F co-ordinated by three Mg-atoms and the Al-content of the mineral. Papin et al.

(1997) investigated the environment of OH-groups in the octahedral sheets of Al-rich

phlogopite and observed a strong preference of OH for a co-ordination by Mg2Al on

the neighbouring cation sites over a co-ordination by three Mg-atoms.

This observation has been confirmed by Fechtelkord et al. (2003a) for synthetic

phlogopites of various Al- and F-contents using 1H, 29Si and 19F MAS NMR

spectroscopy. Moreover, the opposite trend has been found for F which favours pure

Mg-environments. These authors also investigated the Al-content of phlogopite in

relationship to the F-content of the initial starting composition and found a

destabilizing effect on Al-rich phlogopites by F (Fechtelkord et al., 2003a,b).

However, only phlogopites synthesised at 800 °C and 2 kbar have been investigated.

Circone et al. (1991) and Circone and Navrotsky (1992) investigated the

incorporation of Al into the tetrahedral and octahedral sheets by 29Si and 27Al MAS

NMR spectroscopy of synthetic phlogopites synthesised at different temperatures

and pressures. These authors were also the first to combine their experimental data

with computational modelling of cation ordering in the tetrahedral sheets. They

suggested an increased ordering of Si/[4]Al with increasing Al-content of the

phlogopites due to avoidance of Al-O-Al linkages according to Loewenstein’s rule

(Loewenstein, 1954). The influence of fluorine on the observed ordering pattern has

not been considered.

All of these studies have in common that the ordering in both sheets of phlogopite

has been studied separately. However, a relationship between the ordering pattern of

both the tetrahedral and the octahedral sheets is conceivable.

Recently, efforts have been made to describe order/disorder phenomena in

layered silicate structures by means of computational methods. Palin et al. (2001)

1. Introduction

4

first demonstrated the effectiveness of the so-called ‘J-formalism’ in combination with

Monte Carlo (MC) simulations to shed light on Si/[4]Al ordering in the tetrahedral

sheets of muscovite, K Al2 (AlSi3O10) (OH)2. In 2003, these authors extended their

studies to tetrahedral (Si/[4]Al) and octahedral ([6]Al/Mg) ordering in phengite

(K (Al1.5Mg0.5) (Al0.5Si3.5O10) (OH)2). Indeed, a coupling between the ordering in both

sheets was observed, with two [6]Mg-atoms and two [4]Al-atoms forming small clusters

within the structure. Experimental data confirming these results are still missing.

To my knowledge, this study is the first one combining experimental and

computational efforts to obtain an overall picture of the cation and anion distribution

in tetrahedral and octahedral sheets of synthetic Al-rich phlogopite.

29Si, 19F, 1H, and 27Al MAS NMR spectroscopic experiments have been carried

out to gather information on the ordering schemes in both sheets separately.

Moreover, the amount of additional Al incorporated into the phlogopite structure was

estimated from the 29Si MAS NMR spectra to find the maximum of Al-content in

dependence of the F-content of the initial oxide mixture. In contrast to previous

studies, a large number of compositions have been analysed, ranging from hydroxyl-

phlogopites to F-rich compositions and from Al-free to extremely Al-rich starting

mixtures. A synthesis temperature of 600 °C and a pressure of 2 kbar have been

chosen as to complement the data reported by Fechtelkord et al. (2003a,b).

Moreover, {1H/19F} → 29Si CPMAS/HETCOR spectra were recorded to allow an

investigation of the ordering of Mg/[6]Al and OH/F in the octahedral sheet coupled to

that of Si/[4]Al in the tetrahedral sheet. 17O MAS and MQMAS NMR spectroscopic

experiments were performed to check the validation of Loewenstein’s rule. Powder

X-ray diffraction patterns have been analysed to obtain additional information on the

structure on a larger scale. The changes of lattice parameters with increasing Al-

content and the polytypes formed during synthesis have been studied.

The experimental investigations were completed by theoretical calculations

allowing for a deeper understanding of the ordering mechanisms showing in the NMR

spectroscopic results. The calculations were based on the ‘J-formalism’ used by Palin

et al. (2001, 2003) which describes ordering in terms of exchange reactions between

neighboured sites. For each couple of sites the interaction parameter has been

determined and used in Monte Carlo (MC) simulations in order to generate atomic

configurations of lowest energy showing possible ordering patterns. OH-phlogopites

1. Introduction

5

of all Al-contents have been investigated, covering the whole range between

phlogopite (K Mg3 (AlSi3O10) (OH)2) and ‘eastonite’ (K (Mg2Al) (Al2Si2O10) (OH)2). In

this way, it was possible to even study very Al-rich compositions that have not been

accessible experimentally.

2.1. Phlogopite structure and mineralogy

7

2. Theory

2.1. Phlogopite structure and mineralogy

2.1.1 General chemical composition of micas

Phlogopite is a trioctahedral 2:1 layer silicate and belongs to the group of the true

micas. Only a brief introduction to the mica chemistry will be given here. For the

classification and detailed information on the chemistry of true and brittle micas the

reader is referred to Tischendorf et al. (2007) and Rieder et al. (1998).

The general composition formula of the micas can be expressed as

A1 X2-3 (T4O10) (OH)2 (2.1)

In true micas, A is a monovalent cation like K+, Na+, Rb+, Cs+ and (NH4)+. K+ micas

are by far the most common true micas (Tischendorf et al., 2007). Examples for this

type of micas are phlogopite, annite, celadonite, muscovite and polylithionite. A more

abundant true non-K mica is paragonite with Na+ as A-cation. The brittle micas are

defined as having divalent cations like Ca2+ and Ba2+ on the A position. Typical

examples are clintonite and margarite.

A wider range of cations may be incorporated into the X position in the octahedral

layer: Ti, [6]Al, [6]Fe3+, Mn3+, Cr3+, V2+, Fe2+, Mn2+, Mg2+ and Li+. A separation can be

made between dioctahedral and trioctahedral micas. In trioctahedral micas all three

octahedral sites per half unit-cell are occupied (usually by divalent cations), whereas

in dioctahedral micas only ⅔ of these sites are occupied (usually by trivalent cations),

leaving one vacancy for reasons of charge-balancing.

The tetrahedral sheet is mainly occupied by Si and [4]Al, but [4]Fe3+, B and Be may

also be incorporated. For true mica end-members the ratio of tetrahedral Al to Si is

1:3, for the brittle micas this ratio increases to 1:1. OH-groups are mostly found as

anions, but they may also be replaced by F, O, Cl or S.

Phlogopite (K Mg3 (AlSi3O10) (OH)2) belongs to the group of trioctahedral K micas

with Mg being the only cation occupying the X position. Its dioctahedral counterpart is

muscovite (K Al2 (AlSi3O10) (OH)2) with which it does not form a straight solid solution

series (Green, 1981; Robert, 1976). Green (1981) suggested complete solid solution

2. Theory

8

between trioctahedral and ‘2.5-octahedral’ micas while a large immiscibility gap

should exist between the intermediate compositions and dioctahedral micas. Natural

phlogopites often contain considerable amounts of iron and aluminium. There is

complete solid solution between phlogopite and annite (K Fe2+3 (AlSi3O10) (OH)2)

(Wones and Eugster, 1965; Müller, 1972; Wones, 1972). To some extent substitution

of Fe3+ and [6]Al for Mg and [4]Al for Si may occur, with compositions ranging towards

siderophyllite (K (Fe2+2.5Al0.5) (Al1.5Si2.5O10) (OH)2) and ‘K Fe3+

2 (AlSi3O10) (OH)2)’

(Rutherford, 1973; Hewitt and Wones, 1975). Intermediate compositions of the four

end members phlogopite, annite, siderophyllite and the hypothetical composition

‘K Fe2 (AlSi3O10) (OH)2’ are called biotite. Hewitt and Wones (1975) found an upper

substitution limit of Al in synthetic phlogopite corresponding to a composition of

K (Mg2.38Al0.62) (Al1.62Si2.38O10) (OH)2. A phlogopite sample of composition

K (Mg2.08Al0.92) (Al1.92Si2.08O10) (OH)2 has been synthesised by Circone et al. (1991).

Few natural phlogopites also exhibit larger amounts of Mn in the octahedral

sheets. More common is the presence of Fe3+ and/or Ti4+ in the tetrahedral sheet. K

is rarely substituted by Cs, Rb, or Ba. Extensive replacement of OH by F has been

observed for some Fe-poor compositions. For further details see review of

Tischendorf et al. (2007).

Due to experimental restrictions only Fe-free samples have been investigated in

this study. The large diamagnetic effect of iron in the phlogopite crystal structure

leads to broad and featureless lineshapes in the NMR spectra, and no structural

information can be obtained from these samples. Thus, only Al-incorporation via

Tschermak’s substitution and OH ↔ F exchange have been considered.

Al-incorporation is described by a solution of x ‘eastonite’

(K (Mg2Al) (Al2Si2O10) (OH,F)2, a hypothetical end-member) in phlogopite. The

replacement of OH-groups by F is illustrated by the variable y with 0.0 ≤ y ≤ 2.0. The

resulting nominal composition of all samples is then given by

K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y.

2.1.2. The phlogopite structure

Phlogopite, as all micas, is a 2:1 layer silicate: Its structure is formed by layer

packages consisting of two tetrahedral layers and one octahedral layer (see Figure

2.1). The octahedral layer is formed by XO4(OH)2-octahedra that share edges and

form a two-dimensional infinite layer. These layers are sandwiched by two layers of

2.1. Phlogopite structure and mineralogy

9

ab

c

K

Mg

SiAl

O

OF

2:1 layerpackage

interlayer

tetrahedrallayer

octahedrallayer

tetrahedrallayer

OOH/F

K+

Si/Al

Mg/Al

Figure 2.1. View on the stacking sequence of phlogopite-2M1. The unit cell is outlined. After Hendricks and

Jefferson (1939).

2. Theory

10

a

bc

Mg/Al

OH/F

O

K+

Si/Al

Figure 2.2. View on the octahedral sheet of phlogopite-2M1 (after Hendricks and Jefferson, 1939).

TO4-tetrahedra, in which each tetrahedron shares corners with three other

tetrahedra. The fourth oxygen atom (the so-called apical oxygen) is bonded to the

octahedral layer. These layer packages are stacked and separated by the interlayer

cation A.

In phlogopite in the narrower sense the octahedral layer is occupied by Mg only,

i.e. each Mg has six other Mg-atoms as next-nearest-neighbours, and every OH-

group is coordinated by three Mg-atoms in the octahedral layer (see Figure 2.2). As

Al is introduced into the octahedral layer, a second and third environment for OH

being surrounded by two Mg and one Al or one Mg and two Al, respectively, should

be observed theoretically. Figure 2.3 shows that in the tetrahedral layer, each Si-

atom is surrounded by three neighbours in the next tetrahedra, which can be either Si

or Al. Thus, four possible configurations can be expected: Si – Si3, Si – Si2Al, Si –

SiAl2, Si – Al3.

2.1.3. Polytypism

As for all micas, phlogopite may crystallise in different polytypes. The origin of

polytypism in micas – and phyllosilicates in general - is the existence of different

possibilities for stacking the layer packages along the c-axis. The stacking directions

can be best described as a displacement of the OH-groups in adjacent layer

packages.

2.1. Phlogopite structure and mineralogy

11

a

bc

Mg/Al

OH/F

O

K+

Si/Al

Figure 2.3. View on the tetrahedral sheet of phlogopite-2M1 (after Hendricks and Jefferson, 1939). In the right half

of the picture the K+-ions were omitted to show the position of the OH/F site.

Figure 2.4 shows a sketch of a single OH-group within a hexagonal ring of

oxygen atoms belonging to the tetrahedral sheet below. Due to the hexagonal or

ditrigonal symmetry of the tetrahedral sheets there are six possible positions for the

OH-group in the adjacent sheet (shown as dotted circles). These displacements are

described by the six vectors 1, 2, 3, -1, -2, and -3, placed at 60° to each other.

A sequence of different vectors involves a rotation of tetrahedral sheets in

adjacent layer packages against each other, and thus, the different polytypes may be

described by means of the occurring rotations within the maximum degree of

ordering (MDO) theory (Ferraris and Ivaldi, 2002; Nespolo and Ďurovič, 2002).

Polytypes belonging to subfamily A are based on 2n x 60° rotations, whereas those

of subfamily B exhibit only (2n+1) x 60° rotations.

Polytypes of subfamily A are 1M (n = 0), 2M1 (n = 1 and 2) and 3T (n = 1 or 2).

The ideal space group symmetries of these polytypes are C2/m, C2/c and P31,212,

respectively. The octahedral sites in two adjacent layer packages are displaced by

±a/3. The polytypes 2O (n = 1, Ccmm) and 2M2 (n = 0 and 2, C2/c) belong to

subfamily B, in which the octahedral sheets are displaced by ±b/3. Another

hypothetical polytype, 6H (n = 0 or 2, P61,522), is also part of subfamily B. Crystal

structures of the five naturally occurring polytypes are shown in Figure 2.5.

2. Theory

12

OHO

1M

2M1 3T

1

1

-2

1

2

3

1 2

3

Figure 2.4. Schematic illustration of different ways of stacking in micas leading to a different position of the OH-

groups.

Following Ramsdell (1947) the polytype symbols are written as NSn. N is the

number of layer packages in the unit cell. S describes the crystal structure symmetry:

A = triclinic (anorthic), M = monoclinic, O = orthorhombic, Q = tetragonal (quadratic),

T = trigonal, R = rhombohedral, H = hexagonal, and C = cubic. n often denotes the

order in which polytypes of the same symmetry have been discovered.

2.1. Phlogopite structure and mineralogy

13

The most common polytypes are that of subfamily A. Most reports of phlogopite

structure refinements can be found for polytype 1M (e.g., Schingaro et al., 2001;

Alietti et al., 1995), far less publications can be found for the 2M1 and the 3T polytype

(Pini et al., 2008; Fregola et al., 2009; Bigi et al., 1993; Bigi and Brigatti, 1994).

Usually, the 2M1-polytype structure is typical for dioctahedral micas like muscovite.

Ferraris et al. (2001) reported a fluoro-phlogopite-2O, and this polytype has also

been obtained synthetically by Sunagawa et al. (1968) and Endo (1968).

1M 2M1

2M2

3T

2O

Figure 2.5. Crystal structures of the five naturally occurring polytypes in micas (Ferraris and Ivaldi, 2002, p.129).

The 2M1- and the 3T-polytypes have been observed in structures with a high

degree of disorder and long-period stacking orders. Moreover, the distortion of

octahedra has been found to be higher in the 2M1-polytype than in the 1M-polytype,

while there was no relationship between the formation of the 2M1-polytype and the

2. Theory

14

crystal chemistry of the mica (Bigi et al., 1993). As far as the 3T-polytype is

concerned, Fregola et al. (2009) suggested an increase of the stability of this

polytype with increasing K- and Na-concentration in the surrounding fluid during

crystallisation.

2.1.4. Ordering of tetrahedral cations

Since the number of crystallographically independent sites in the tetrahedral

sheets of phlogopite is restricted due to symmetric considerations, the possible

cationic ordering patterns are different for the three polytypes discussed in the

previous chapter. Only one tetrahedral site is present in the 1M-polytype of space

group C2/m, and no ordering of Al and Si on specific sites is possible (Ferraris and

Ivaldi, 2002). In the 2M1-polytype and the 3T-polytype with space group C2/c and

P3112, respectively, two independent tetrahedral sites may be distinguished and thus

cation ordering may occur.

The presence of tetrahedral cation ordering in micas is difficult to clarify using

standard X-ray diffraction techniques. As aluminium and silicon both have a similar

scattering factor, a refinement of tetrahedral site occupancies is not possible. In his

review on cationic ordering in micas, Bailey (1975, 1984a) investigated tetrahedral

long-range ordering of Al and Si using a statistical analysis of small deviations from

the mean T-O bond lengths in the tetrahedral sheets reported in the literature. He

concluded that three factors stabilise long-range ordering in the tetrahedral sheets:

(1) the stacking arrangement of the 3T-polytype, (2) a tetrahedral composition of

Si/Al close to 1:1, and (3) a “phengitic” composition (i.e. muscovite of celadonitic

composition). The most common mica minerals, however, are found to be

disordered: muscovite-2M1, phlogopite-1M, and biotite-1M.

Loewenstein’s rule states that on exchange of Si by Al in the tetrahedral sheets,

occupancy of directly neighboured tetrahedral sites by two Al-atoms is avoided

(Loewenstein, 1954). This was confirmed by Lipsicas et al. (1984) for several mica-

type compositions. These authors proposed the existence of Al-Si short-range

ordering in the tetrahedral sheets. Herrero et al. (1985a,b, 1987) further brought

forward the argument that this type of ordering should be governed by the

minimisation of local charge imbalance due to the substitution Si ↔ Al. Following

their model of homogeneous dispersion of charges (HDC), charges should be

distributed evenly in the tetrahedral sheet at low Al-contents. One example for the

2.1. Phlogopite structure and mineralogy

15

consequences resulting from this type of ordering is that at least one AlO4-

tetrahedron will be found in each of the hexagonal rings of tetrahedra.

Closer to a composition of Si/[4]Al = 1:1, as found in margarite, long-range

ordering should be established due to a coupling between adjacent hexagonal rings,

with Al and Si occupying the tetrahedral sites alternately. Circone et al. (1991)

confirmed these conclusions for hydroxyl-phlogopites of different Al-contents. These

authors emphasised the non-existence of Al-O-Al linkages in phlogopite. However,

Loewenstein’s rule may not be valid in disordered high-temperature structures, and

several authors (e.g., Langer et al., 1981) claimed to have evidence of Al-O-Al

linkages in micas due to specific infra-red absorption bands.

2.1.5. Ordering of octahedral cations

Octahedral cation ordering is possible for all three phlogopite polytypes (Brigatti

and Guggenheim, 2002). Three distinct octahedral sites are present in the 3T-

polytype with space group P3112 (Ferraris and Ivaldi, 2002). Two of them, M2 and

M3, are cis-coordinated by OH, while the third, M1, is trans-coordinated (Brigatti and

Guggenheim, 2002). The latter is usually vacant in dioctahedral micas. In the 1M-

and 2M1-polytypes (space group C2/m and C2/c, respectively) the two cis-

coordinated sites M2 and M3 are symmetrically related by a mirror plane, and thus,

only two octahedral sites may be distinguished, M1 and M2 (Ferraris and Ivaldi,

2002).

Based on refinements of octahedral site occupancies the micas may be divided

into three subgroups (Ďurovič, 1994, Nespolo and Ďurovič, 2002, Brigatti and

Guggenheim, 2002, Ferraris and Ivaldi, 2002, Mercier et al., 2005): I. Homo-

octahedral micas in which all three octahedral sites are occupied by the same kind of

cation or the same statistical average of different kinds of cations, II. meso-

octahedral micas in which two of the octahedral sites are occupied by one cation and

the third by a different one, and III. hetero-octahedral micas in which each of the

three octahedral sites is occupied by a different cation. This classification is

independent from a division into several types of micas based on differences in the

size of the octahedral sites, i.e. the average cation-anion bond lengths of the different

sites as proposed by Weiss et al. (1992). Mercier et al. (2005) and Brigatti and

Guggenheim (2002) emphasised that even for structures having equal average

occupancies for all three octahedral sites the size of the octahedra may be different

2. Theory

16

and vice versa. Therefore, the investigation of ordering in the octahedral sheets of

micas via X-ray diffraction techniques must take into account differences in the

average cation-anion bond lengths as well as a refinement of octahedral-site

occupancies.

Toraya (1981) argued that larger cations of lower charge should be ordered on

the M1 site to ensure minimum cation-cation repulsion between neighbouring

octahedra. Indeed, Cruciani and Zanazzi (1994) found a preferential partitioning of

highly charged cations like Al3+ on the M2 site. Also, as Fe2+ enters the octahedral

sheet following the phlogopite-annite join, there are usually differences in the site

occupancy factors and the size of the octahedra, with Fe2+ slightly preferring the M1

site while the tendency for Al3+ to occupy the M2 site is increased (Brigatti et al.,

2000). Nevertheless, it is not uncommon for all three sites to be equal in size and to

show the same site occupancies in 1M-polytype micas (Brigatti and Guggenheim,

2002).

According to Brigatti and Guggenheim (2002) there have been several reports of

micas with intermediate compositions between dioctahedral and trioctahedral micas.

However, these are most likely to be mixtures of one dioctahedral and one

trioctahedral phase. In dioctahedral micas the vacant site is usually the M1 site, so

this should also be the case for any intermediate compositions.

In their study, Mercier et al. (2005) found that natural and synthetic 1M-polytype

single crystals were geometric meso-octahedral and always showed some degree of

ordering cations into octahedral sites of different size. In contrast the authors showed

that synthetic powder samples were always of the homo-octahedral type, and the

cations were statistically distributed in the octahedral sheet. They concluded that the

synthetic powder samples had not reached the equilibrium state yet, which was

present in the single crystals due to a much longer crystallisation time. The small

difference in configuration energies between the homo- and the hetero-octahedral

structure indicates that geometrical reasons cannot be the driving force of a potential

ordering in tri-octahedral micas.

Papin et al. (1997) showed in their IR spectroscopic investigation of phlogopite

that Mg and Al are not distributed statistically in the octahedral sheet. It has been

found that F prefers a co-ordination by Mg only, whereas OH-groups favour an

2.1. Phlogopite structure and mineralogy

17

environment containing also Al. These results have been confirmed by Fechtelkord et

al. (2003a) using 1H and 19F MAS NMR spectroscopy.

It should also be noted that partial ordering of the octahedral sites is not

uncommon in micas in general (Bailey, 1984a). An example would be atoms of type

A occupying one of the two M2 sites, and the same atom type A being randomly

distributed on the other M2 and the M1 sites together with atom type B.

2.1.6. Exchange of OH by F in the octahedral sheet

In the basic mica structure octahedral cations are co-ordinated by four oxygen

atoms and two hydroxyl-groups. Theoretically, OH may easily be replaced by F, as

both anions exhibit the same charge of -1 and a similar anionic radius in a threefold

co-ordination of 1.34 and 1.30 Å, respectively, (Shannon, 1976). Nevertheless, OH-

substitution by F is limited for most mica compositions. Synthesis of pure-F

trioctahedral and dioctahedral micas is possible in water-free systems. On addition of

water, however, F strongly prefers a trioctahedral over a dioctahedral environment

(Robert et al., 1993; Papin et al., 1997; Boukili et al., 2001).

In a trioctahedral sheet like that of phlogopite, all apical oxygens of the

tetrahedral sheet are well balanced, and the interaction between the proton of the

hydroxyl-group and these oxygen atoms is very low. The O-H vector is directed

vertically to the sheets, away from the three positively charged Mg-ions, and pointing

towards the interlayer cation K+. The strong repulsion between the like charges leads

Figure 2.6. Sketch of two rings of tetrahedra belonging to adjacent layer packages. In between, the interlayer

cation K+ is shown. In dioctahedral micas the proton of the OH-group is pointing into the vacancy, minimizing the

repulsion between like-charged proton and K+. (Brigatti and Guggenheim, 2002, p.41)

2. Theory

18

to a widening of the structure with larger distances between the single layer

packages. In this situation, it is highly advantageous to substitute OH by F and at the

same time replace the H+-K+ repulsion by a K+-F- attraction.

The opposite is true for dioctahedral micas where the surrounding tetrahedral

apical oxygen atoms are strongly underbonded due to the vacancy. The hydroxyl-

group may no longer be regarded as an entity but rather acts as a dipole with the

proton being directed away from the two occupied sites and pointing towards the

vacancy (Figure 2.6). This leads to a much lower H+-K+ repulsion while the

interaction between proton and apical oxygen atoms is increased. OH ↔ F exchange

is less favourable because F is not able to contribute to the local charge balancing in

the way OH does.

These effects also have a strong influence on the thermal stability of phlogopite:

Wones (1967) found a decomposing temperature of less than 905 °C at 100 bar for

pure hydroxyl-phlogopite, whereas melting temperatures of 1345 – 1390 °C (at 1

kbar) have been reported for fluoro-phlogopite (Van Valkenburg and Pike, 1952;

Shell and Ivey, 1969).

2.1.7. Phlogopite mineralogy

Micas in general occur in a wide range of rocks. They can be found in intrusive

and extrusive igneous rocks and in upper mantle rocks as well as in metamorphic

rocks formed over a wide range of pressure and temperature conditions.

The main occurrence for phlogopite is in contact-metamorphosed limestones,

dolomites and ultrabasic rocks. These rocks often show low iron contents, so that

nearly pure Mg-phlogopites can be found (e.g., Schreyer et. al, 1980). The mineral is

also often found in kimberlites in India, South Africa and Canada (e.g., Rao et al.,

2009; Zurevinski et al., 2008), and thus, thought to be present in considerable

amounts in the upper mantle. In these depths, phlogopite and other micas are

supposed to play a key role as a carrier of volatiles which can be released through

complex reactions and then change the melting conditions of the surrounding rocks

(Virgo and Popp, 2000).

Phlogopite is also common in intrusive igneous rocks of granitic compositions,

especially in those of late magmatic stages, where F can be strongly enriched

(Carroll and Webster, 1994). In these rocks it often contains higher amounts of iron,

2.1. Phlogopite structure and mineralogy

19

ranging to biotite in composition, and large crystals up to several meters in size can

be formed.

In metapelitic rocks of nearly all temperature and pressure ranges, phlogopite

may occur as an accessory mineral. However, more often biotite is formed instead

due to the higher iron content of these rocks. Other phlogopite-bearing rocks are

contact-metamorphic calc-silicate rocks and marbles.

2. Theory

20

2.2 Solid-state NMR spectroscopy

Solid-state nuclear magnetic resonance (NMR) spectroscopy is a useful tool for

structural investigations which probes the local environment of the atoms in the

lattice up to the second co-ordination sphere. Hence, it is a complementary method

to other techniques like X-ray diffraction (XRD) and infra-red (IR) spectroscopy.

Information which can be obtained in NMR spectroscopic experiments contains

the co-ordination number and type of co-ordinating atoms, bond angles, and dynamic

processes in the lattice. In contrast to X-ray diffraction experiments, NMR

spectroscopy is able to easily detect the position and local environment of protons in

the lattice, and even amorphous structures can be investigated.

NMR spectroscopy is especially useful for the investigation of micas like

phlogopite. These materials often cause severe problems in the analysis of their

X-ray diffraction patterns: The small crystallite size of synthetic samples gives rise to

broad reflections, and the plate-like shape makes it necessary to integrate models in

the fitting procedures which deal with the preferred orientation of crystals during the

experiments. Moreover, these structures are often of low symmetry, mostly

monoclinic, leading to a large number of reflections which makes it difficult to

distinguish between the mica reflections and those of impurity phases. Another

problem is caused by stacking faults and mixtures of several different polytypes of

one single phase leading to broad bumps in the background of the patterns (Chapter

4.7).

Therefore, NMR spectroscopy is an ideal tool to investigate the cation and anion

distribution in the layers of phlogopite. Nuclei which can be measured are 1H, 19F, 27Al and 29Si, and – if enriched in the sample - 17O. Another potential nucleus is 25Mg.

However, 25Mg NMR experiments are not as easy to perform given the low

magnetogyric ratio of this nucleus.

The next chapters will give a short introduction to the basics of the NMR

spectroscopic experiments used to investigate the atomic arrangement in the sheets

of phlogopite. The interactions influencing the spectra as well as techniques based

on these will be explained briefly. For more detailed information, the reader is

referred to textbooks on solid-state NMR spectroscopy like those of Abragam (2007),

Ernst et al. (1994), and Slichter (1996).

2.2 Solid-state NMR spectroscopy

21

2.2.1. Interactions influencing NMR lineshapes

2.2.1.1. Zeeman interaction

NMR spectroscopic experiments are based on the interaction between the

magnetic part of a radio-frequency wave and the magnetic moment of a specific type

of nucleus in the structure.

The magnetic moment results from the nuclear spin I, thus, all nuclei with I > 0

possess a magnetic moment which is connected to the nuclear spin I by

Iµ ˆˆ , 2h

(2.2)

with being the magnetogyric ratio of the specific nucleus and h being Planck’s

constant. The energy levels corresponding to the (2I+1) eigenstates are degenerated

unless brought into an external magnetic field 0B

. In this case, the energy levels split

up. This effect is called Zeeman interaction, and the corresponding Hamiltonian is

defined as

00ˆˆˆ BIBH zz

. (2.3)

zI is the z-component of the nuclear spin operator which interacts with the external

magnetic field 0B

. Solving the Schrödinger equation for the Zeeman interaction, one

obtains the corresponding energy eigenvalues

0BmEm

(2.4)

where m is the magnetic quantum number.

In NMR spectroscopic experiments, an electromagnetic wave is irradiated to

induce transitions between the energy levels. These transitions are only allowed for

∆m = ±1 (selection rule) and the energy of the incoming wave must be equal to the

difference in energy between the two energy levels, ∆E:

00 BE . (2.5)

The resonance frequency 0 is called the Larmor precession frequency of the

specific nucleus.

2. Theory

22

2.2.1.2. Chemical shift interaction

The chemical shift interaction is caused by the influence of the electron shell on

the nucleus. The external magnetic field induces a motion of electrons in the electron

shell. This movement in return induces a secondary magnetic field which changes

the resulting magnetic field at the nucleus. It may either enhance the static magnetic

field leading to a higher effective magnetic field at the nucleus, the so-called de-

shielding effect. The external magnetic field may also be lowered by this secondary

magnetic field, and thus, the nucleus be shielded.

The result is a change in the frequency of the electromagnetic wave necessary

for the excitement of transitions between energy levels and a shift of the signal in the

spectrum: The signal position is moved to higher frequencies for de-shielding effects

and to lower frequencies for shielding effects. The secondary magnetic field depends

on the electron density distribution around the nucleus which in return is influenced

by the chemical environment of the atom. Each separate environment will therefore

lead to a different signal in the resulting NMR spectrum.

The chemical shift Hamiltonian CSH is defined as

0ˆˆ BIH zCS

(2.6)

where is the chemical shielding tensor. This tensor is first described in the so-

called principle axes system (PAS) with the origin placed in the centre of the nucleus.

However, for a simpler mathematical treatment the tensor is then transformed to the

laboratory axes system (LS), with the z-direction of the LS being equal to the z-

direction of the external magnetic field.

The chemical shielding tensor is set up in the PAS in a way that the strongest

absolute interaction is directed along the z-direction:

isoyyisoxxisozz

zz

yy

xx

,

00

00

00

(2.7)

with iso being the isotropic chemical shift which is defined by the trace of the

symmetric chemical shielding tensor.

2.2 Solid-state NMR spectroscopy

23

Triso 3

1 . (2.8)

Another parameter influencing the chemical shift lineshape of an NMR signal is

the chemical shift anisotropy aniso which is defined as follows (Duer, 2002):

)( isozzaniso . (2.9)

The asymmetry parameter describes the deviation from axial symmetry

isozz

xxyy

(2.10)

and ranges between 0 ≤ ≤ 1. Thus, the chemical shift anisotropy lineshape can

give information on the symmetry of the co-ordination polyhedron around a specific

nucleus.

As the Larmor-frequency is dependent of the external magnetic field, the exact

signal position also changes with increasing field strength. This makes it difficult to

compare spectra recorded at different spectrometers unless the field strength is

exactly the same. Therefore, the chemical shift is given relative to a reference

material:

ppmref

refxiso

610

. (2.11)

x is the resonance frequency of the observed nucleus, and ref is the resonance

frequency of the reference material.

2.2.1.3. Dipolar interaction

Another interaction influencing the magnetic field at the observed nucleus is the

interaction with other nuclei possessing a magnetic moment nearby, the so-called

dipolar interaction. These nuclei can be either of the same type of nucleus, i.e.,

homo-nuclear dipolar interaction, or the interaction can take place with a different

type of nucleus, i.e., hetero-nuclear interaction.

In the laboratory frame the Hamiltonian for the homo-nuclear dipolar interaction

between two atoms i and j with nuclear spin operators iI and jI can be written as:

2. Theory

24

jijz

iz

ij

iDD IIII

rH ˆˆˆˆ1cos3

2

1

4ˆ 2

3

20

. (2.12)

0 is the permeability of vacuum, and r is the distance between the two nuclei i and j.

is the angle between the internuclear vector r

and the direction of the external

magnetic field 0B

(Figure 2.7.). Corresponding to that the Hamiltonian for the hetero-

nuclear dipolar interaction between atoms i and j can be described as

jiz

ij

jiDD SI

rH ˆˆ21cos3

2

1

4ˆ 2

30

, (2.13)

where jS is the nuclear spin operator of the nucleus not being under

investigation in the experiment.

B0

rij

Îi

Îj

Figure 2.7. The dipolar interaction between two spins i and j.

In the case of a multi-spin system in polycrystalline solids like the phlogopite

samples investigated in this study, the dipolar interaction leads to a significant

broadening of the observed NMR signals, and the typical features resulting from

chemical shift interaction can not be distinguished anymore.

Such signal shapes can be described by a Gaussian frequency distribution )(g

with normalised area:

2.2 Solid-state NMR spectroscopy

25

2

20

)(

)(2ln12ln

)(

eg . (2.14)

0 is the frequency of the maximum of function )(g , and is the full width at half

maximum. However, in most cases the lineshape is best described by the so-called

second moment 2M , which is given by the mean quadratic linewidth.

dgM )()( 202 . (2.15)

In the special case of a pure Gaussian lineshape the second moment can be

calculated directly from the full width at have maximum:

2ln2

2

2

M . (2.16)

In return, from the second moment 2M , information on the structural arrangement

can be obtained using the van Vleck equation (van Vleck, 1948) which is valid for

polycrystalline materials in which all environments for the observed nucleus i are on

average the same throughout the whole structure:

2222 ISIIM . (2.17)

rij and rik are the average internuclear distances between interacting nuclei, 2II is

the homo-nuclear second moment and described as

624

2

02 1)1(

45

3

ijiiiII r

II

. (2.18)

2IS is the hetero-nuclear second moment and can be written as

6222

2

02 1)1(

415

4

ikkkkiIS r

II

. (2.19)

This connection between the second moment and the direct environment of the

nucleus allows the estimation of distances between atoms in cross-polarisation

experiments discussed in Chapter 2.2.2.2.

2. Theory

26

As phlogopite contains many nuclei with magnetic moments and some of them

even with a natural abundance of the specific nucleus of 100 % (19F, 27Al) or a value

close to that (1H), dipolar interaction of both types must be considered for every

experiment. This means that all experiments have been performed using the magic

angle spinning technique (Chapter 2.2.2.1.) to average out dipolar interactions and to

decrease the line-broadening caused by this interaction.

2.2.1.4. Quadrupolar interaction

For nuclei with a nuclear spin of I > 1/2, another interaction must be taken into

account, the quadrupolar interaction. This is due to the electric charges not being

distributed equally in the nucleus but in form of a quadrupol leading to an electric

quadrupolar moment Q. In case of a non-spherical distribution of electric charges in

the surrounding electron shell, an electric field gradient (EFG) ikV is generated with

which the quadrupolar moment may interact.

As a result the difference between the energy levels is not equal for all levels

anymore (Figure 2.8). However, in case of a first-order perturbation, when the

quadrupolar interaction is low in comparison to the Zeeman interaction, the energy

level difference of the central transition (1/2 → -1/2) remains unchanged. The

quadrupolar Hamiltonian QH may then be written as

N

i

ii

i

ii

iQ IVI

II

eQH

1

ˆˆ)1(2

ˆ

. (2.20)

After insertion of the Wigner matrixes this equation becomes

2cossin

2

1

2

1cos3)1(ˆ3

)12(4ˆ 2

22

2

IIIII

qQeH z

iiQ

(2.21)

with the two Euler angles and . The principal elements of the EFG are then

defined as

,

00

00

00

zz

yy

xx

V

V

V

V xxyyzz VVV . (2.22)

2.2 Solid-state NMR spectroscopy

27

m = 3/2

m = 1/2

m = -1/2

m = -3/2

L

L

L

L 2+ (L

L 1(

Zeemaninteraction

first-orderquadrupolarinteraction

second-orderquadrupolarinteraction

Figure 2.8. Schematic illustration of the changes of the differences between the energy levels for Zeeman, first-

order and second-order quadrupolar interaction for a spin 3/2 nucleus (after Medek et al., 1998).

The definitions of the quadrupolar coupling constant QC , the quadrupolar frequency

Q and the asymmetry parameter are as follows:

,2

h

qQeCQ zzVeq ; (2.23)

hII

qQeQ )12(4

2

; (2.24)

,zz

xxyy

V

VV 10 . (2.25)

If the quadrupolar interaction is comparable or even larger than the Zeeman

interaction, second-order terms must be taken into account to describe the

interaction accurately. In this case, even the energy level difference of the central

transition is not the same anymore leading to a quadrupolar shift QS additional to

the chemical shift CS mentioned above. The quadrupolar shift is defined as follows:

2

22

2

0 3

11

)12(

3)1(9)1(

40

3

II

mmIICQQS ,

2

00 . (2.26)

2. Theory

28

0 is the Larmor frequency of the nucleus, and m is the magnetic quantum number.

The second-order quadrupolar interaction also leads to a significant broadening of

the signal for the central transition.

When investigating phlogopites, the quadrupolar interaction has to be considered

for 27Al and 17O which possess a nuclear spin of I = 5/2. As for the dipolar interaction,

magic-angle spinning (Chapter 2.2.2.1) has been performed to average out first-order

quadrupolar interaction. However, this is not possible for the case of second-order

quadrupolar interaction. The existence of several possible transitions has been used

for multiple quantum magic angle spinning experiments (Chapter 2.2.2.4).

2.2.2. Experimental techniques

2.2.2.1. Magic angle spinning NMR spectroscopy

Magic angle spinning (MAS) is a technique which is routinely used to average out

the anisotropic parts of all first-order interactions in order to narrow the spectral

lineshapes and to increase the spectral resolution. To achieve this, the sample has to

be spun fast about an axis with an angle of = 54°44’ related to the external

magnetic field.

It can be shown (Maricq and Waugh, 1979) that each interaction operator

consists of three parts:

)(ˆˆˆˆ,, tHHHH statiso . (2.27)

isoH ,ˆ is an isotropic part only present in the case of chemical shift interaction.

TCH isoˆˆ

, (2.28)

C is a constant depending on the interaction, T contains the spin operator, and is

the corresponding interaction tensor. The second term statH ,ˆ is a time-dependent

angular term:

)1cos3(2

1

3

2ˆˆ 2,

TCH stat

SSS 2cossin

2)1cos3(

2

1 22 . (2.29)

2.2 Solid-state NMR spectroscopy

29

statH ,ˆ becomes zero if the spinning axis is oriented with an angle of 54°44’ to the z-

axis of the external magnetic field due to the term )1cos3( 2 . The last term )(ˆ tH is

a time-dependent term:

tStCtStCTCtH rrrr 2sin2cossincos

3

2ˆ)(ˆ2211 (2.30)

nC and nS are time-independent trigonometric terms (Maricq and Waugh, 1979). If

the rotation frequency r is much larger than the spectral frequency width of the

static signal, )(ˆ tH is time-averaged and only the isotropic signal will be present in

the resulting spectrum. If the rotation frequency r is smaller than , so-called

spinning sidebands will appear in the spectrum at distances equal to the rotation

frequency. These are marked by asterisks in the spectra discussed in the results

section.

2.2.2.2. Cross-polarisation magic angle spinning NMR spectroscopy

As has been mentioned earlier, the dipolar interaction between neighbouring

nuclei can be used to gather information on the distance between these nuclei and

on atomic arrangements in the structure.

After Hartmann and Hahn (1962) and Pines et al. (1972, 1973) first reported this

technique it has been excessively used to increase the low spectrum intensity of

nuclei with a low magnetogyric ratio and/or a low natural abundance. For this

1(H) H 1(H) Si 1(Si) 1(Si) = B = B =

0(H) H = B0 0(Si) Si = B0

1H 29Si

Figure 2.9. Sketch of the energy levels of 1H (‘cold spin revervoir’, left) and 29Si (‘hot spin reservoir’, right). A

transfer of energy from the hot system to the cold one is only possible if the Hartmann-Hahn-condition is fulfilled

(middle).

2. Theory

30

purpose, magnetisation is transferred from another nucleus with a high natural

abundance and a high magnetogyric ratio to this specific nucleus via dipolar

interaction. In the framework of this study, 29Si has been chosen to accept

magnetisation which has been transferred from either the 1H or the 19F nucleus.

However, it was not the purpose to increase signal intensity for 29 Si, but to use the

dipolar interaction to gather information on the distance between these nuclei and on

atomic arrangements in the structure. In this chapter the basics of this technique will

be explained shortly for {1H} → 29Si cross-polarisation magic angle spinning

(CPMAS) NMR spectroscopy, but the method works for {19F} → 29Si CPMAS NMR

alike.

To allow the energy transfer between both nuclei, the radiofrequency field 1B has

to be adjusted in such a way that the distance in energy between the energy levels is

the same in both systems S and I, and the so-called Hartmann-Hahn condition is

fulfilled (Hartmann and Hahn, 1962):

)29(129)1(11 SiSiHH BB . (2.31)

Thermodynamically, this process can be described as a heat exchange between a

cold and a hot spin reservoir (Figure 2.9). The distribution of spins in the two systems

is a Boltzmann distribution, and the fractional number Ni/N of spins occupying a

certain energy level Ei is given by

i

Tk

E

Tk

E

i

B

i

B

i

e

e

N

N

, (2.32)

where Bk is the Boltzmann constant. If the energy difference is low, many spins

already occupy the excited state and, thus, cannot undergo a transition any more,

leading to low signal intensities. However, if the spin temperature of the system and

thus, the energy difference is increased more spins occupy the ground state and are

available for excitation.

A two-pulse sequence has to be applied for the magnetisation transfer, as shown

in Figure 2.10a. The first pulse in the 1H channel brings the 1H magnetisation into the

xy-plane. This is followed by a long spin-lock pulse to hold the magnetisation there,

2.2 Solid-state NMR spectroscopy

31

while a 90°x pulse for 29Si starts the transfer of magnetisation. The duration of this

90°x pulse is called the contact time t. After the systems have been brought into

contact, acquisition starts.

In contact-time dependent CPMAS NMR experiments the contact time is increased

stepwise in a number of experiments and the amount of magnetisation transferred is

recorded. Two cases must be considered: that of large 1H spin reservoir in the 29Si

environment and that of an isolated spin system (Walther et al., 1990).

90°x

1H

29Si

tcontact time acquisition

90°x

high power Pya)

90°x

1H

29Si

tcontact time acquisition

90°x

high power Pyb)

evolution

decoupling

decoupling

Figure 2.10. Pulse sequence schemes for {1H} → 29Si CPMAS (a) and 2D {1H} → 29Si HETCOR (b) NMR

experiments.

2. Theory

32

M(T )+M(T )1 HSi

M(T )1

M(T )HSi

Figure 2.11. Example of a magnetisation function for the case of a large proton spin reservoir. The curve has

been calculated according to equation (2.33) using the following parameters: M0 = 4*1010 a.u., T1ρ = 45 ms, and

THSi = 9 ms.

In the first case, the magnetisation curve is a sum of two parts:

)(1

)( 10 HSiT

tT

t

eeM

tM

,

1T

THSi . (2.33)

HSiT is the cross polarisation time, and 1T is the spin-lattice relaxation time in the

rotating frame. One part is made up by the exponential increase of magnetisation at

the 29Si nucleus with increasing contact time until saturation is reached and the

magnetisation is constant (Figure 2.11). This increase is a function of the cross

polarisation time HSiT and depends on the number of protons in the Si-environment

and the distance between these protons and the 29Si nucleus, HSir . Meanwhile,

magnetisation is back-transferred to the 1H nucleus again, as this looses energy due

to spin-lattice relaxation. The resulting decay in magnetisation depends on the spin-

lattice relaxation time in the rotating frame 1T and thus, on the number of vibrational

or rotational processes in the lattice.

2.2 Solid-state NMR spectroscopy

33

Figure 2.12. Example of a magnetisation function for the case of an isolated spin system. The curve has been

calculated according to equation (2.34) using the following parameters: M0 = 4*1010 a.u., T1ρ = 12 s, THSi = 6 ms,

THH = 50 ms, a = 0, and b = 500.

In the case of an isolated spin system with distinct H-Si distances additional

oscillatory contributions have to be considered and equation (2.33) changes to

bteeaaee

T

TM

tM HHHHHSi T

t

T

t

T

tT

t

HSi 2

1cos

2

1

2

1)1(

1)( 2

3

1

0 1

(2.34)

with HHT being the spin diffusion time and a ranging from 0 ≤ a ≤ 1. b is the

oscillation frequency and given by

3

20 )1cos3(

42

1

HSi

SiH

rb

. (2.35)

A typical plot of the curve is shown in Figure 2.12.

These relationships have been used to estimate the shortest distance between

protons and 29Si nuclei. In a first step, the contact-time dependent curves have been

fitted by least-squares procedure according to (2.33), and the cross polarisation time

HSiT has been determined.

2. Theory

34

According to Pines et al. (1973), the cross polarisation time is connected to the

homo-nuclear second moment 2HH and the hetero-nuclear second moment

2HSi in the following way:

2

21

HH

HSi

HSiHSi

CT

. (2.36)

HSiC is a geometrical constant for which as an approximation the value for the similar

kaolinite structure has been taken:

2

HSiC . (2.37)

(Hayashi and Akiba, 1994). 2HH has been calculated according to equation (2.18)

after a structural model of Tateyama et al. (1974), considering only the five shortest

given H-H distances. This allowed for the calculation of 2HSi from equation (2.36)

and the calculation of the mean H-Si distance according to equation (2.19).

2.2.2.3. 2D hetero-nuclear correlation CPMAS NMR experiments

A variation of the CPMAS NMR experiment described above, the 2D hetero-

nuclear correlation (HETCOR) CPMAS experiment, can be used to gather

information on the atomic arrangement in the octahedral sheets of phlogopite related

to that in the tetrahedral sheets. Information on the proton environment can be

transferred together with the magnetisation and correlated to the corresponding 29Si

environment.

In this experiment, the contact time is kept constant. Instead, an additional delay

is inserted between the first pulse in the 1H channel and the contact pulse, and the

length of this delay is increased in a number of experiments (Figure 2.10b). In this

way, the amount of magnetisation being transferred changes as the proton

magnetisation decreases and information on the proton environment can be detected

indirectly.

The result is a two-dimensional spectrum giving connectivity information: The

more often two specific environments are located next to each other in the structure,

the higher the corresponding signal.

2.2 Solid-state NMR spectroscopy

35

2.2.2.4. Multiple quantum MAS NMR spectroscopy

It has already been discussed earlier (Chapter 2.2.1.4) that first- and second-

order quadrupolar interactions lead to a significant broadening of signals in spectra of

nuclei with spin I > ½. For the investigation of phlogopite this causes problems for 27Al (I = 5/2) and 17O (I = 5/2) NMR spectra, making it difficult to distinguish between

signals of different environments in the phlogopite structure as well as in impurity

phases. Therefore, multiple quantum magic angle spinning (MQMAS) experiments

have been performed to obtain additional information on the lineshape of the

observed NMR signals.

Spinning the sample rapidly about an axis has been routinely used in the MAS

technique (Chapter 2.2.2.1) to average out first-order interaction. Then the evolution

of a spin coherence after excitation can be described by multirank expansion of its

phase :

tPmCmttm SQCS )(cos)(2),,( 000

tPmCtPmC SQSQ )(cos)(),()(cos)(),( 444222 . (2.38)

CS is the isotropic chemical shift, and and are the Euler angles describing the

orientation of the principle axes system of the quadrupolar tensor to the external

magnetic field. Q0 is the zero-rank, i.e. isotropic, quadrupolar shift, ),(2 Q is a

second-rank anisotropic frequency similar to the chemical shielding and dipolar

interaction anisotropies, and ),(4 Q is a fourth-rank anisotropic frequency resulting

from the fact that second-order effects are proportional to the interaction squared.

20 3)1(2)( mSSmmCS , (2.39)

312)1(82)( 22 mSSmmC S , (2.40)

534)1(182)( 24 mSSmmC S

(2.41)

are zero-, second- and fourth-rank coefficients depending on the spin S and the order

m of the transition. )(cos nP are Légendre polynomials of cos :

1)(cos0 P ; (2.42)

2. Theory

36

2

1cos3)(cos

2

2

P ; (2.43)

8

3cos30cos35)(cos

24

4

P . (2.44)

Now it becomes obvious that it is impossible to average out all anisotropic terms

of equation (2.38) by spinning about just one angle. First approaches to get rid of the

line-broadening anisotropic parts have been dynamic angle spinning (DAS) and

double rotation (DOR). In the DAS method the sample is spun about two different

angles alternately (Mueller et al., 1990), while the DOR equipment consists of a

smaller rotor placed in an outer rotor, and both rotors are let spun simultaneously to

fulfil both angle conditions (Samoson et al., 1988).

However, both methods are technically complex, which is why for this study the

MQMAS technique (Frydman and Harwood, 1995; Medek et al., 1995) has been

chosen, in which the spinning angle , usually the magic angle, is fixed. Instead,

spins are allowed to evolve during times 1t and 2t under the effect of two transitions

1m and 2m , chosen in such a way that the following second-order averaging

conditions are fulfilled:

0)()( 222112 tmCtmC SS, (2.45)

0)()( 224114 tmCtmC SS. (2.46)

In the upper part of Figure 2.13, the pulse scheme used for most of the MQMAS

experiments performed in the framework of this study is shown. The lower part of the

same figure shows the corresponding coherence scheme with the coherence

pathways. At the beginning of the experiment the coherence order p is always zero.

During the first pulse – the preparation pulse - all possible transitions with coherence

orders p = +5, +4, +3, …, -4, -5 are excited because the 1m selection rule is

not valid any more due to perturbations of the quadrupolar interaction.

1m has been chosen to be 3/2 which means that the corresponding transitions

with coherence orders p = +3 and p = -3 have to be filtered. This can be easily

done due to the fact that upon shifting the excitation pulse of n -quantum coherences

by degrees the resulting signals will undergo a phase shift of n .

2.2 Solid-state NMR spectroscopy

37

3 2 1 0-1-2-3

p

preparation pulse

conversionpulse

t1 t2

-4-5

4 5

1 2 R

Figure 2.13. Top: Pulse scheme for the 27Al 3QMAS NMR experiment. Bottom: Corresponding coherence path

scheme.

Applying the appropriate pulse sequence will lead to a summation of the desired

resonances, in the case of 27Al 3QMAS the triple-quantum resonances, and will make

all other transitions with different coherence order vanish.

The value of 2m can only be ½, as only single-quantum excitations are visible in

NMR, and thus the second pulse has to convert the triple quantum excitations in a

detectable single quantum resonance with coherence order p = -1. This conversion

pulse always possesses a phase of = 0°. Following Frydman and Harwood

(1995), the whole six scan phase cycle applied in the two-pulse MQMAS experiment

is:

1 = 0°, 60°, 120°, 180°, 240°, 300°

2 = 0°, 0°, 0°, 0°, 0°, 0° (2.47)

R = 0°, 180°, 0°, 180°, 0°, 180°.

2. Theory

38

However, the resulting signal ),( 21 ttS X is a linear combination of the echo and anti-

echo signals, where the echo pathway is p = 0 → 3 → -1 and the anti-echo pathway

is p = 0 → -3 → -1. To obtain a pure absorption mode real spectrum, these two

signals have to be separated which can be achieved by shifting the first pulse by

90°/ p , where p is the order of the coherence in the 1t evolution period, in this case

3 (Massiot et al., 1996). The new phase cycle is then:

1 = 30°, 90°, 150°, 210°, 270°, 330°

2 = 0°, 0°, 0°, 0°, 0°, 0° (2.48)

R = 0°, 180°, 0°, 180°, 0°, 180°

and a second signal ),( 21 ttSY is generated. The echo and anti-echo signals,

),( 21 ttS E and ),( 21 ttS A respectively, can then be calculated from

),(),(),( 212121 ttiSttSttS YXE (2.49)

),(),(),( 212121 ttiSttSttS YXA (2.50)

Afterwards, a shearing transformation is necessary to obtain the isotropic spectrum,

and a 1t -dependent first-order phase correction has to be applied:

).()','(' 21),(

2121 tSetS E

tiE

(2.51)

).()','(' 21),(

2121 tSetS A

tiA

(2.52)

where

122,1 12

19)( tt (2.53)

(Massiot et al., 1996). The pure absorption mode 2D spectrum ),( 21 S is obtained

by Fourier-transforming both signals with respect to 1't and then combining in the

following way:

)','(')','(')','( 212121 AE SSS . (2.54)

The position of the signals (in ppm) in the isotropic projection is then given by

2.2 Solid-state NMR spectroscopy

39

1

393

108

31

17 2

20

26

Qiso

(2.55)

for I = 5/2. is the difference between the isotropic chemical shift and the

reference, 0 is the Larmor-frequency of the nucleus, and QC and are the

quadrupolar coupling parameters. Q is the Zeeman-frequency and defined as:

)12(2

6

II

CQQ

. (2.56)

Equation (2.55) is also a useful tool to check the goodness of the fit of a 27Al

MAS NMR spectrum by calculating the theoretical signal position in the isotropic

dimension from the fitted quadrupolar parameters and then comparing this value to

the signal position found in the 27Al MQMAS NMR spectra.

2. Theory

40

2.3 J-formalism and Monte-Carlo simulations

The combination of the so-called ‘J-formalism’ and Monte Carlo simulations used

for this study has been first reported by Bosenick et al. (2001) and Warren et al.

(2001), and since then the method has been applied to cation ordering in various

mineral systems like spinels (Palin and Harrison, 2007) and pyroxenes (Warren et

al., 2001). It has been shown to be a very effective technique to study cation ordering

in layered structures through investigation of the tetrahedral sheets of muscovite

(Palin et al., 2001), the tetrahedral and octahedral sheets of phengite (Palin et al.,

2003) and the octahedral sheets of minerals of the illite/smectite-group (Sainz-Díaz

et al., 2003a,b; Palin and Dove, 2004). However, these have all been dioctahedral

phyllosilicates whereas the phlogopite structure studied here is trioctahedral.

The J-formalism describes the energy of a given cation configuration as a sum of

separate pair interactions among the ordering cations. A set of pair interaction

parameters, the ‘Js’, is generated from empirical interatomic potentials and lattice

energy minimisation methods using the General Utility Lattice Program (GULP)

(Gale, 1997). These parameters are then used in Monte Carlo simulations of the

cation ordering as a function of temperature. This chapter will present a short outline

of the computation techniques.

For the simulations, the phlogopite structure is considered a network of sites on

which atoms may order, and at first only the tetrahedral sheet has been studied.

The energy for the ordering of the two atom types Si and Al on the network sites

of the tetrahedral sheet can then be described as a sum over all interactions:

)(0n

SiAln

SiAln

SiSin

SiSin

AlAln

nAlAl ENENENEE , (2.57)

where nN are the numbers of pairs of kind n . It can be shown that this relationship

may be simplified to

)2('0n

SiAln

nSiSi

nAlAl

nAlAl EEENEE (2.58)

(Bosenick et al., 2001).

2.3 J-formalism and Monte-Carlo simulations

41

a

b

J1

J2

J3

J4K+

O

OH/F

Mg/Al

Si/Al

Figure 2.14. Assignment of J -parameters within one tetrahedral sheet.

This means that the total energy can be expressed as the exchange energy of two

like pairs (Al-Al and Si-Si) by two unlike pairs (Al-Si) to form Al-O-Si bonds instead of

Al-O-Al and Si-O-Si bonds. This exchange energy is called J so that equation (2.58)

becomes

nn

nAlAl JNEE '0 . (2.59)

If nJ is positive, it is energetically favourable for neighbouring sites to contain

different types of atoms, and if nJ is negative, same atoms are preferred on

neighbouring sites. It should be emphasised that the formation of an Al-Al pair always

implies the formation of a Si-Si pair, i.e., Loewenstein’s rule of Al-O-Al linkage

avoidance (Loewenstein, 1954) is at the same time a rule of Si-O-Si avoidance, if

both types of atoms order on one type of site.

2. Theory

42

a

b

K

Mg

SiAl

O

OF

OOH/F

K+

Si/Al

Mg/Al

Figure 2.15 Assignment of octahedral J -parameters.

Interaction pairs are defined by the distance between the two corresponding

sites. At first, J -parameters for the interactions within a single tetrahedral sheet

have been defined (Figure 2.14). 1J describes the interaction between two directly

neighboured tetrahedra, 2J that of second-nearest tetrahedra, 3J that of third-

nearest tetrahedra, and so on. In the same way interaction parameters present within

a single octahedral sheet have been defined (Figure 2.15). Additionally, for an

investigation of a relationship between the ordering in the different sheets, J -

parameters for sites in directly neighboured sheets have been assigned (Figure

2.16). These are active for adjacent octahedral and tetrahedral sheets, for two

tetrahedral sheets of the same layer package, and for tetrahedral sheets of different

layer packages. All defined interaction parameters iJ and the values determined with

GULP are given in Table 2.1.

2.3 J-formalism and Monte-Carlo simulations

43

a

c

K+

O

OH/F

Si/Al

Mg/Al

Figure 2.16. Examples of tetrahedral intralayer (green), tetrahedral interlayer (blue), and octahedral-tetrahedral

(red) J -parameters.

2. Theory

44

Table 2.1. Definition of the 16 J -parameters for phlogopite and their values in eV averaged over runs for the two

compositions x = 0.0 and x = 1.0. The approximate error bar is 0.05 eV.

Ji Kind of interaction Value

[eV]

J1

J2

J3

J4

Tetrahedral intrasheet

1.072

0.262

0.132

0.091

J5

J6

J7

Tetrahedral intralayer

0.196

0.146

0.109

J8

J9

J10

J11

Tetrahedral interlayer

0.020

-0.009

0.003

-0.029

J12

J13

J14

J15

Octahedral intrasheet

0.576

0.136

0.151

0.059

J16

J17

J18

J19

Tetrahedral-octahedral

interaction

-0.607

-0.368

-0.226

-0.170

Once all J s have been defined, a large number of configurations with random

distribution of atoms on the sites are created, and the energy of each configuration is

computed following the minimisation of lattice energy with the program GULP (Gale,

1997). The cell has been a 2 x 1 x 1 supercell of the monoclinic unit cell

(10.6 Å x 9.2 Å x 10.1 Å). This has been done for two compositions with x = 0.0 and

x = 1.0, respectively, to test the dependence of the J s on composition. It has been

found that there seems to be no relationship between the J s and the composition.

2.3 J-formalism and Monte-Carlo simulations

45

For the intrasheet J s, the values achieved with GULP have been tested against a

first-principles fitting with the SIESTA code (Ordejón et al., 1996). The values

obtained with the SIESTA code were slightly smaller than those of the GULP fitting;

except for 4J . However, both values follow the same trends of relative magnitude.

With these J -parameters Monte-Carlo (MC), simulations have been

performed, using the programme Ossia99 (Warren et al., 2001) to obtain the atomic

arrangements of lowest energy for 0.0 ≤ x ≤ 1.0. This code uses an implementation of

the density functional theory (DFT) which is based on a real-space description of the

electron density. The size of the calculation scales linearly with the system size

making possible a description of large and complex systems. Moreover, Ossia99

uses the pseudo-potential method taking into account only valence electrons while

the inner electrons are represented by pseudo-potentials.

The model Hamiltonian has been derived from the Ising-model, a

mathematical model for the description of ferromagnetic structures which can be

easily transferred to the ordering of atoms on sites (Warren et al., 2001). An ordering

variable jS describing the occupancy of site j is defined in such a way that 1jS if

the site is occupied by Si, and 0jS if this site is occupied by Al. Then, the sum over

all interactions is formed whereby every interaction is only counted once, and the

corresponding Hamiltonian can be written as

ij

jjjiij SSSJH ˆ (2.60)

The last term is a chemical potential which operates if the A atoms prefer specific

sites in the structure, while the first term is associated with the bonds.

A random distribution of atoms is then applied to the supercell, and a random

change is proposed which results in an energy change of EEE , with the aim

to find the configuration of lowest energy. The change is accepted, if the energy of

the new state is lower than that of the actual state. However, it must also be allowed

for the system to accept a new state of higher energy in order to find the global

minimum instead of only a local minimum. Thus, a probability function is applied in

this case:

2. Theory

46

Tk

E

BeEEEp

)( (2.61)

where Bk is the Boltzmann constant and T is the temperature of the system. For the

first steps, the temperature is set to very high values of 3000 K to allow atomic

movement in the structure. In a process of ‘simulated annealing’ the temperature is

stepwise lowered to about 100 K. This makes large changes in energy more unlikely,

and the movement of atoms is ‘frozen’.

More than 800M steps have been performed for each run, and for each

composition five runs have been simulated for better statistics. Most runs have used

a supercell of 16 x 16 x 2 of the phlogopite unit-cell containing 7168 active sites on

which ordering may occur. However, also 24 x 24 x 2 supercells have been used for

comparison. As expected, the latter showed a smaller dispersion in the values for the

occupancy of specific sites, but the ordering patterns were similar for both types of

supercells.

All simulations have been carried out by Dr. Alberto García Arribas, Institut de

Ciència de Materials de Barcelona, CSIC, Bellaterra, Spain, and Dr. Javier López-

Solano, Universidad del Pais Vasco, Bilbao, Spain.

3.1. General approach

47

3. Experimental and analytical methods

3.1. General approach

A number of 85 phlogopite samples have been synthesised and investigated

using NMR spectroscopy and powder X-ray diffraction techniques (Figure 3.1).

Fine-grained oxide powders have been prepared by the gelling method according

to Hamilton and Henderson (1968), and from these, phlogopite samples have been

synthesised via hydrothermal synthesis in cold seal bombs at T = 600 °C and

p = 2 kbar. Following the incorporation of aluminium into the structure via

Tschermak’s substitution and the exchange of OH-groups by fluorine, the nominal

composition of the reagent powder was

K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F(2-y). (3.1)

The nominal Al-content of the samples is denoted as xnom. As not all aluminium

was incorporated into the phlogopite structure, this value must not be mixed up with

the real Al-content of the phlogopites, which can be estimated using 29Si MAS NMR

spectroscopy (see Chapter 4.2.1) and is therefore named xest.

Figure 3.2 gives an overview on the composition of the synthesised samples. The

majority of samples have been prepared at the Ruhr-Universität Bochum. Some

additional samples synthesised at the same conditions have been prepared by

Predrag Vulić at the University of Innsbruck and kindly provided for analysis in the

frame of this thesis. Samples synthesised at a temperature of 800 °C and a pressure

of 2 kbar, have been provided by Dr. Michael Fechtelkord (Ruhr-Universität Bochum).

The preparation of these samples has been described in Fechtelkord et al. (2003a).

Two more samples with compositions xnom = 0.6, y = 0.5 and xnom = 0.8, y = 0.5,

respectively, have been prepared at a temperature of 800 °C to check the

dependence of the starting materials used on the formation of impurity phases (see

Chapter 3.2 for details).

Only 64 of the 85 samples synthesised at 600 °C have been used for further

analysis. The samples with low Al- and high OH-contents had to be discarded due to

problems with the crystallisation of forsterite during burning the gel over the Bunsen

flame. Due to their high Si-content amorphous oxide mixtures of this composition

3. Experimental and analytical methods

48

KNO (aq)3 Mg(NO ) (aq)3 2 Al(NO ) (aq)3 3 TEOS KF(aq) / NH F(aq)4

+NH (aq) Sol-Gel3

heating up to 800°C

sealing in gold capsules

hydrothermal synthesis runT = 600°C, p = 2 kbar

NMR experiments

29

19

Si MAS

F MAS

Al MAS

H MAS

( O MAS)

27

1

17

{ H} -> Si CPMAS

{ F} -> Si CPMAS

Al MQMAS

( O MQMAS)

1 29

19 29

27

17

powder XRD

characterisation

+ethanol + water

Figure 3.1. Flow chart of synthesis and characterisation of phlogopite samples.

3.1. General approach

49

hypothetical end-membercomposition ‚eastonite‘ K(Mg Al )(Al Si O )(OH) F2 1 2 2 10 y 2-y

samples synthesised and describedby Fechtelkord et al. (2003a, b), synthesis temperature 800 °C

samples synthesised at 600 °C andnot used for analysis

samples synthesised at 600 °C andcharacterised in the framework of this study

compositions for which samples have been remade using O enriched water17

Figure 3.2. Nominal compositions K(Mg3-xAlx)(Al1+xSi3-xO10)(OH)yF2-y of the initial oxide mixtures used for the

synthesis of phlogopite samples.

seem difficult to prepare. Other samples were not considered because it must be

assumed that the gold capsules leaked during synthesis.

The samples synthesised at 600 °C have been characterised routinely using 29Si, 1H, 19F and 27Al MAS NMR to estimate the real aluminium content of the phlogopites

and gather information on the different types of ordering in the structure. Impurity

phases have been identified using the NMR spectra as well as XRD powder patterns.

More sophisticated NMR spectroscopic experiments like {1H} → 29Si and

{19F} → 29Si CPMAS/HETCOR MAS, and 27Al MQMAS experiments have been

carried out only on a limited number of samples to gather additional structural

information.

Three samples with composition xnom = 0.5 and y = 0.5, 1.0, and 1.8 have been

synthesised using water with 28 or 75.0 – 80.9 at% 17O (Sigma-Aldrich). On these 17O MAS and 17O MQMAS NMR experiments have been performed to check

Loewenstein’s rule of avoidance of Al-O-Al linkages in the structure (Loewenstein,

1954).

3. Experimental and analytical methods

50

3.2 Sample preparation

3.2.1. Preparation of gels

In order to obtain highly reactive and homogenous starting materials for the

hydrothermal synthesis of phlogopite samples, the method of Hamilton and

Henderson (1968) has been used to prepare amorphous, fine-grained oxide mixtures

(‘gels’). In these, a network of amorphous SiO2 forms a ‘gel’ in which the other

reagents are held.

1M aqueous solutions of KNO3 (VWR International), Mg(NO3)2*6H2O (Sigma-

Aldrich), Al(NO3)3*9H2O (Sigma-Aldrich), KF (VWR International), and NH4F (Merck)

have been prepared to add potassium, magnesium, aluminium and fluorine. For

hydroxyl-rich compositions (y ≥ 1) both potassium and fluorine have been added in

form of KF, and KNO3 has been used to add more potassium if necessary. For

fluorine-rich compositions (y < 1), potassium has been added in form of KNO3, and

NH4F has been used for fluorine to avoid excess potassium in the mixture.

The aqueous solutions have been titrated in Teflon containers. Liquid tetraethyl

orthosilicate (TEOS, Merck) has been used as source of silicon. TEOS has also been

titrated considering its density at room temperature. Afterwards ethanol has been

added in order to obtain a homogenous solution of nitrates and TEOS.

To start the gelling process, a solution of ammonium hydroxide (33%) has been

added until the solution was alkaline and a mixture of metal oxides and hydroxides

precipitated. The mixture has then been dried in a drying-cabinet that was heated

stepwise from 50 to 120 °C. Afterwards the resulting hard pellets have been broken

and heated up to 800 °C in a platinum crucible over a Bunsen flame to evaporate

water, ethanol, ammonia and nitrogen oxides. The powder has been ground in an

agate mortar and used for hydrothermal synthesis.

3.2.2. Hydrothermal synthesis

For synthesis the oxide mixtures have been sealed in gold tubes with

approximately 9 wt% of water. Typically, an amount of 150 to 300 mg of powder has

been used. The outer diameter of the gold tubes was 4 mm, and the thickness of the

gold was 0.5 mm. After sealing, the capsules had a length of 4 to 5 cm.

3.1. General approach

51

Syntheses were carried out with a conventional hydrothermal apparatus with

horizontal Tuttle-type pressure vessels (Tuttle, 1949; Luth and Tuttle, 1963; Figure

3.3).The cold-seal vessels were externally heated, and water was used as pressure

medium. After insertion of the gold capsules, the bombs were filled with Ni-rods in

order to prevent water circulation in the bomb. The temperature has been measured

using Ni/NiCr-thermocouples which were inserted into a borehole in the bomb,

parallel to the gold capsule.

At the beginning of the experiment, pressure and temperature were increased

simultaneously and then set to the final values of T = 600/800 °C and p = 2 kbar. It

can be assumed that run conditions were stable after 1-2 hours run duration. The

pressure has been corrected several times during the experiment, if necessary. After

one week the furnaces were opened and temperature and pressure decreased

simultaneously.

With the long capsules and the temperature being measured outside the vessel’s

interior, the estimated error in temperature is ± 20 °C. The error in pressure set is ±

50 bars.

Ni/NiCr-thermocouple

gold capsulewith sample Ni-rod

cold seal bomb

to pressuresystem

~10 cm

cap

Figure 3.3. Schematic illustration of the cold seal bombs used for the synthesis of phlogopite.

After the run the bombs were cooled at air with the furnaces opened and the

pressure decreasing slowly with temperature. This procedure usually took about 3

hours. The sample powder was taken out of the gold capsule, dried at 120 °C and

ground in an agate mortar again before characterisation.

3. Experimental and analytical methods

52

3.3. NMR spectroscopic experiments

The majority of NMR experiments have been performed at room temperature on

a Bruker ASX 400 NMR spectrometer using a 89 mm wide-bore magnet with a field-

strength of 9.34 T. The obtained spectra have been fitted using the DMFIT software

developed by Dr. Dominique Massiot (Massiot et al., 2002).

High field MQMAS experiments have been performed by Dr. Ulrike Werner-

Zwanziger, Dr. Josef Zwanziger, and Dr. Michael Fechtelkord at a field strength of

16.45 T on a Bruker Avance NMR spectrometer at Dalhousie University, Halifax, NS,

Canada. Additional experiments have been performed by the group of Dr. Jürgen

Haase on a Bruker Avance NMR spectrometer at a field strength of 17.6 T at the

Department of Interface Physics and the Magnet-Resonanz-Zentrum of the University

of Leipzig.

3.3.1. 1H MAS NMR experiments

1H MAS NMR experiments have been performed at 400.13 MHz with a standard

Bruker 4 mm MAS probe at rotation frequencies of 12.5 kHz. Liquid tetramethylsilane

(TMS) was used as an external standard. A pulse length of 2 µs and a recycle delay

of 10 s were used, and 128 scans were accumulated. The spectral width was

125 kHz. After the measurements, a spectrum of the rotor without sample has been

recorded and subtracted from the original spectrum to eliminate the broad signal

resulting from protons in the rotor and the rotor cap.

3.3.2. 29Si MAS NMR experiments

The 29Si MAS NMR spectra were recorded at an operating frequency of

79.49 MHz with liquid tetramethylsilane (TMS) as external standard. A standard

Bruker 7 mm MAS probe has been used with spinning frequencies of 4 kHz, a single

pulse duration of 4 µs, a recycle delay of 10 s, and a spectral width of 50 kHz. Some

experiments have been repeated with a recycle delay up to 120 s, but no change in

signal intensity has been observed. A number of 7000 to 25000 scans have been

accumulated.

3.3.3. 27Al MAS NMR and 27Al 3QMAS NMR experiments

The 27Al MAS NMR experiments have been performed at 104.268 MHz with a

standard Bruker 4 mm MAS probe at rotation frequencies of 12.5 kHz. Single pulse

3.3. NMR spectroscopic experiments

53

duration was 0.6 µs to ensure homogeneous excitation of the central as well as all

satellite transitions. A recycle delay of 0.1 s was used and a number of 25000 scans

have been accumulated. The spectral width was 125 kHz, and an aqueous solution

of AlCl3 has been used as an external standard.

For the 27Al 3QMAS spectra recorded at 9.34 T, the spectral width has been

reduced to 50 kHz in the F2-dimension. A recycle delay of 0.2 s has been used.

Pulse lengths were 2.5 and 2.0 µs for the excitation and the conversion pulse,

respectively. An initial delay between the two pulses of 3 µs has been chosen which

has then been increased stepwise by 20 µs in a number of 56 or 128 experiments.

For each experiment a number 400 to 1000 scans have been accumulated, and

before each scan a number of 12 dummy scans were performed to ensure complete

saturation. F1-axis labelling has been done following the C3a-convention (Amoureux

and Fernandez, 1998; Millot and Man, 2002).

The experiments at 16.45 T have been recorded at a transmitter frequency of

182.47 MHz, a H-F/C-P probe head and 2.5 mm rotors. Potassium alaun has been

used as a secondary reference with a chemical shift of -0.033 ppm. The 1D 27Al MAS

NMR spectra were acquired with a nominally 9° pulse at 95 kHz rf field strength at

10.0 and 22.0 kHz sample spinning with a repetition time of 500 ms. The 27Al 3QMAS

spectra have been recorded with a three pulse sequence, where the last echo pulse

and its timings allowed for split-t1 mode and whole echo acquisition. The excitation

pulse length was 3.6 µs, the conversion pulse length was 1.2 µs. The selective echo

pulse lasted for 11.0 µs, and a recycle delay of 1s has been chosen. 672 - 2016

scans have been accumulated for each of the 80 to 100 slices. The F1 axis was

scaled and referenced according to the Cz-convention (Millot and Man, 2002) but

inverted due to the echo acquisition.

Additional high-field 27Al MAS NMR experiments have been performed at 17.6 T.

The transmitter frequency was 195.28 MHz, and a 2.5 mm MAS probe has been

used at spinning rates of 25 kHz. The single pulse duration and recycle delay were

0.4 µs and 0.5 s, respectively. 400 scans have been accumulated for each

experiment.

3.3.4. 19F MAS NMR experiments

For 19F MAS NMR experiments a Bruker 4 mm MAS probe has been used. 300

scans were accumulated at a frequency of 376.46 MHz and rotation frequencies of

3. Experimental and analytical methods

54

12.5 kHz. Single pulse duration was 2 µs, and a spectral width of 125 kHz and a

recycle delay of 10 s have been used. As external reference a liquid p-C6H4F2 sample

has been measured, and the frequency with highest signal intensity has been set to

-120 ppm with respect to liquid CFCl3.

3.3.5. 17O MAS and 17O MQMAS NMR experiments

The 17O MAS NMR experiments were performed at a frequency of 54.25 MHz

and a sample spinning speed of 12.5 kHz using a 4 mm Bruker MAS probe. De-

ionised water enriched with 28 at% 17O has been used as reference.

For the 1D 17O NMR experiments, a short pulse length of 0.6 µs has been chosen

in order to ensure equal excitation of all possible transitions, and the recycle delay

has been set to 1 s. A number of 50000 to 70000 scans have been accumulated. The

spectral width was 50 kHz.

For the 17O 3QMAS experiments at 9.34 T, excitation and conversion pulse

lengths of 22.5 µs and 9.5 µs have been used, respectively, and the recycle delay

was 2 s. 900 scans have been accumulated for each experiment, and 128

experiments have been recorded for each spectrum. The initial delay between the

excitation and the reconversion pulse was 3 µs and the delay has been increased in

steps of 40 µs. The F1-axis has been labelled according to the C3a-convention

(Amoureux and Ferandez, 1998; Millot and Man, 2002).

For the 17O 3QMAS NMR experiments performed at 16.45 T, a H-F/C-P probe

head and 4 mm rotors have been used, and the chemical shift was referenced

against water. The 1D 17O MAS NMR spectra have been recorded with pulse length

of 1.6 µs, a repetition time of 1s and spinning speeds of 10.0 and 22.0 kHz. For the 17O MQMAS NMR experiments, excitation and conversion pulse lengths of 10 µs and

2 µs, respectively, have been used. A third echo pulse of 12.8 µs was used for split-

1d mode and whole echo acquisition. For the sample with nominal enrichment of 17O

of 80.5 to 89 at%, a recycle delay of 2s has been chosen, and 192 scans were

accumulated for each of the 128 experiments. The recycle delay for the sample with 17O nominal enrichment of 28 at% was 1 s, and 7200 scans have been acquired for

36 experiments. The F1 axis was scaled according to the Cz-convention (Millot and

Man, 2002) and inverted due to the echo acquisition.

3.3. NMR spectroscopic experiments

55

3.3.6. {1H} → 29Si CPMAS/HETCOR experiments

The {1H} → 29Si CPMAS/HETCOR NMR spectra were recorded at transmitter

frequencies of 400.13 MHz and 79.49 MHz for 1H and 29Si, respectively. A standard

7 mm Bruker MAS probe has been used at rotation frequencies of 4 kHz. The 90°

pulse length for 1H was 7.6 µs (rf(1H) = rf(

29Si) = 33 kHz), and the recycle delay was

5 s. A number of 360 – 400 scans were accumulated for each experiment, and

tetramethylsilane was used as a reference for the chemical shift of both 1H and29Si.

For the CPMAS contact-time dependent experiments contact times of 0.1 to 120 ms

have been chosen. For the 2D cross polarisation (HETCOR) experiment a contact

time of 2 ms and a t1-increment of 20 µs have been used.

3.3.7. {19F} → 29Si CPMAS/HETCOR experiments

The transmitter frequencies for the {19F} → 29Si CPMAS/HETCOR NMR

experiments have been 79.49 MHz and 376.45 MHz for 29Si and 19F, respectively.

Sample spinning speed was 5.8 kHz in a standard 7 mm Bruker MAS probe.

Tetramethylsilane (TMS) and p-C6H4F2 (δ = -120 ppm) have been used as reference

for 29Si and 19F, respectively. The 90° pulse length for 19F was 5.6 µs (rf(19F) =

rf(29Si) = 45 kHz), and a recycle delay of 5 s has been used. For the 1D CPMAS

contact-time dependent experiments contact times of 0.1 to 120 ms have been

chosen. In the 2D CPMAS HETCOR experiment a contact time of 5 ms and a t1-

increment of 10 µs have been used. A number of 64 experiments have been

performed for each spectrum.

3. Experimental and analytical methods

56

3.4. X-ray diffraction experiments

Powder X-ray diffraction (XRD) patterns have been recorded in collaboration with

the group of Prof. Dr. Tonči Balić-Žunić on a Bruker D8 powder diffractometer at the

University of Copenhagen. The samples have been measured in Bragg-Brentano

geometry with Cu Kα1 radiation (λ = 1.5406 Å, 45 kV, 40 mA) and a Ge(111)-crystal

as primary monochromator. A 2θ range of 2 to 70° has been covered, the step size

was 0.02 °2θ and the counting time was 1 second per step. Samples were prepared

on flat Si-sample holders providing a sample thickness of 0.5 mm.

Additional measurements have been performed on a Philips PW1830/40 powder

diffractometer at the Ruhr-Universität Bochum using Cu Kα radiation (λ = 1.5418 Å,

45 kV, 30 mA) and a secondary monochromator. Intensities have been recorded

between 5 and 60 °2θ with a step size of 0.02 °2θ and a counting rate of 2 seconds

per step. Flat glass sample holders have been used for the experiments.

More detailed measurements have been recorded on a Siemens D5000

difractometer in collaboration with Dr. Bernd Marler at the Ruhr-Universität Bochum.

Cu Kα1 radiation (λ = 1.5406 Å, 45 kV, 35 mA) and a Ge(111)-crystal as primary

monochromator have been used. Only a 2θ range of 22-28° has been investigated

because this was the part of the pattern in which satellite reflections had been

observed before. The step size was 0.0078 °2θ, and a position sensitive detector

(PSD) with an opening angle of 6 °2θ has been used. Samples have been prepared

in glass capillaries with 0.03 mm diameter. Temperature-dependent measurements

have been carried out by heating the sample with hot air up to about 500 °C.

3.5. Scanning electron microscopy

57

3.5. Scanning electron microscopy

Scanning electron microscopy (SEM) has been performed to investigate the

crystal shapes and sizes of phlogopite and impurity phases using a LEO Model 1530

(ZEISS SMT) field emission gun scanning electron microscope with an acceleration

voltage of 20 kV.

For the investigations small amounts of sample powder have been brought on

aluminium mounts coated with adhesive conducting carbon tabs. The sample has

been coated with gold and brought into a vacuum chamber, and the SEM images

have been recorded by scanning the sample surface with the high-energy electron

beam in a raster-scan pattern.

When the electron beam hits the sample surface, a number of radiation signals

are generated including secondary electrons (SE) and X-rays. For a detailed analysis

of the surface structure of a sample, the secondary electrons are detected which

result from an ionisation of the k-orbitals of atoms at or near the sample surface and

are usually of low energy (< 50 eV). The number of secondary electrons produced

depends on the angle between the electron beam and the surface, making it possible

to generate images with a three-dimensional appearance.

The X-rays generated may be analysed using an energy dispersive X-ray

detector system (EDX) which allows for the detection of all emitted wavelengths at

once. As these X-rays are element specific, the chemical composition of a small part

of the sample can be determined. However, the chemical data obtained in this way

are of rather low quality compared to electron microprobe analyses because of the

rough surface of the samples.

4.1. General description of samples

59

4. Results and discussion

4.1. General description of samples

A number of 64 phlogopite samples has been prepared having a nominal

composition of

K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F(2-y) (4.1)

with 0.0 ≤ xnom ≤ 1.6 and 0.0 ≤ y ≤ 2.0. Synthesis runs were performed at

temperatures of 600 °C and pressures of 2 kbar and a run duration of one week. As

expected, the obtained powder samples consist mainly of phlogopite (at least 70 %).

The phlogopite crystals exhibit a plate-like shape typical for the mica minerals

with their cleavage along the (001)-plane (see scanning electron microscope (SEM)

images in Figure 4.1). The crystal size is very small, ranging from 0.2 to 1 µm in

diameter. The impurity phases could not be identified directly in these images.

However, analyses of the several µm large clusters shown in Figure 4.1a taken with

an energy-dispersive X-ray (EDX) detector indicate that these are actually crystals of

several impurity phases with much smaller phlogopite platelets sticking to them. The

samples have been used for NMR experiments before taking SEM images so that

pressing the sample into the rotors may have led to this effect.

The low crystallinity of the samples leads to broad NMR signals and XRD peaks.

This is especially a problem for the determination of the chemical composition of the

phlogopites. Electron-microprobe (EMP) analysis was not possible due to the small

crystal size, so that 29Si MAS NMR spectroscopy was the only way to estimate the

real Al-content incorporated into the phlogopite structure.

Several impurity phases have been identified by the corresponding signals in the

NMR spectra and in the powder XRD patterns (see also Chapter 4.7). A list of

phases determined by XRD is given in Table 4.1. Nearly pure phlogopite formed at

low Al-contents (x < 0.2), and the amount of phlogopite in the sample steadily

decreases with increasing amount of Al in the initial oxide mixture. The minimum

fraction of phlogopite determined by XRD was about 70% for Al-rich samples

(0.8 ≤ xnom ≤ 1.6), but this need not necessarily be the true amount of phlogopite in

the mixture. Only crystalline phases will appear in the XRD patterns whereas,

amorphous phases cannot be detected this way.

4. Results and discussion

60

2 µm

a)

b)

c)

Figure 4.1. Scanning electron microscope (SEM) images of typical run products. The samples consist of several

µm large crystals of impurity phases (a) with much smaller phlogopite crystals sticking to them (b). The phlogopite

platelets exhibit a diameter of less than 1 µm and often show a more or less hexagonal shape (c).

4.1. General description of samples

61

Table 4.1. Crystallite sizes and relative amounts of phases in phlogopite samples determined by LeBail-fitting of

the phlogopite XRD patterns.

x y crystallite size [nm]

R-value

Phl [%]a)

Crn [%]a)

Chl [%]a)

Sel [%]a)

0.8 0.0 77 (4) 6.12 70 23 7

0.1 0.2 47 (2) 6.44 99 1

0.0 0.5 138 (6) 6.22 100

0.2 0.5 122 (5) 5.99 99 1

0.5 0.5 73 (4) 5.97 98 2

0.7 0.5 62 (3) 6.68 86 14

0.0 1.0 80 (5) 6.24 100

0.5 1.0 30 (1) 6.29 100

0.6 1.0 46 (3) 5.99 100

0.8 1.0 46 (3) 6.10 100

0.7 1.2 63 (3) 5.96 100

0.2 1.6 41 (2) 6.12 90 10

0.4 1.6 43 (2) 6.02 100

0.5 1.6 42 (2) 5.86 95 5

1.2 1.6 55 (2) 6.02 82 18

0.5 2.0 43 (2) 6.07 97 3

1.6 2.0 57 (2) 6.07 67 33

a) The error bar of the relative amounts is ± 5 %. Phl = phlogopite, Crn = corundum, Chl = chlorite, Sel = sellaite.

Indeed, several additional signals have been observed in the 19F MAS NMR

spectra of F-rich samples (y < 1.0) making up relative signal intensities of up to 45%

(see Table 6.4 in Appendix). These might be caused by amorphous components like

starting material which has not reacted during synthesis. It should be noted that the

relative intensities of these signals may be overestimated due to short spin-lattice

relaxation times of these compounds.

The most prominent impurity phase, however, was aluminium oxide (Al2O3). It

has already been observed in previous studies (e.g., Circone et al., 1991;

Fechtelkord et al., 2003a,b) and could be identified by the corresponding signal in the 27Al MAS NMR spectra in virtually all samples with xnom ≥ 0.2, independent from the

OH/F-ratio of the sample. The amount of Al2O3 formed in the samples cannot be

estimated from these spectra because of signal overlapping and the quadrupolar

nature of the 27Al nucleus (see Chapter 4.5 for details). XRD patterns are less

sensitive to small amounts of impurity phases, and for this reason corundum is not

4. Results and discussion

62

mentioned for most of the samples listed in Table 4.1. Nevertheless, the amount of

Al2O3 can roughly be estimated by analyzing the patterns of high-Al samples. The

most Al-rich sample (xnom = 1.6, y = 2.0) contains about 30% Al2O3, for samples

containing slightly less Al this is in the range of 15 – 25%. Moreover, 27Al MAS and

MQMAS spectra revealed that Al2O3 has not completely crystallised as corundum (α-

Al2O3) as has been observed by Fechtelkord et al. (2003a,b), but is a probably some

disordered pre-phase resulting from lower synthesis temperatures.

Fechtelkord et al. (2003a,b) also reported the formation of potassium aluminium

hexafluoride, K3AlF6*0.5H2O, for their F-rich (y = 0.5) samples synthesised at 800 °C.

This is confirmed by the samples investigated in the frame of this study in which

potassium aluminium hexafluoride is present for all samples of y = 0.5 and xnom ≥ 0.6

(Table 6.4 in Appendix). However, this impurity phase has also been observed in

samples of intermediate (y = 1.0, xnom = 0.7) and very low (y = 1.8, xnom ≥ 1.2)

F-contents. In the latter, it makes up about 30% of the whole signal intensity in the 19F MAS NMR spectrum. Although this value is highly overestimated due to the fast

spin-lattice relaxation of the AlF63- complex, considerable amounts of potassium

aluminium hexafluoride must be present in these samples. A corresponding signal in

the 27Al MAS NMR spectra was difficult to observe due to signal overlapping.

Compared to the samples synthesised at 800 °C, the amount of potassium

aluminium hexafluoride formed in these samples is rather low (Chapter 4.3.3.2).

A second F-rich impurity phase is MgF2. It could be identified in the 19F NMR

spectra and in the XRD pattern of the sample with x = 0.8 and y = 0.0.

Some of the samples still contained ammonia, which indicates that not all of the

nitrogen has been removed from the gel before synthesis. However, the maximum

relative intensity of this signal in the 1H MAS NMR spectra was 2%, so it can be

assumed that this does not influence the phlogopite synthesis.

Kalsilite (KAlSiO4) has been observed by Fechtelkord et al. (2003a,b), but this

phase was never present in the samples synthesised at 600 °C. No corresponding

signal was present in the tetrahedral region in the 27Al MAS NMR spectra, and that

has been checked with 27Al MQMAS NMR experiments. These authors used KF

instead of NH4F for adding fluorine to the initial gel mixture which could have lead to

a formation of the K-rich impurity phase kalsilite. However, kalsilite was also present

4.1. General description of samples

63

in two samples synthesised in the framework of this thesis but at 800 °C. It can be

concluded that this impurity phase only forms at elevated temperatures.

4. Results and discussion

64

4.2. Ordering of cations in the tetrahedral sheets of phlogopite

4.2.1. Samples of intermediate to low F-contents (1.0 ≤ y ≤1.8)

In 29Si MAS NMR experiments the local environment of Si-atoms and the

ordering of Si/[4]Al in the tetrahedral sheet can be probed. Moreover, 29Si MAS NMR

spectroscopy is the most precise method to determine the amount of Al incorporated

into the structure during synthesis.

The 29Si MAS NMR spectra of the phlogopite samples show up to five signals

(Figure 4.2). Four of them, positioned at at -80, -83, -87, and -90 ppm, result from

different Si environments in the phlogopite structure (Weiss et al., 1987). An

additional signal due to an impurity phase has been found in the 29Si MAS NMR

spectra at about -94 to -95 ppm for some of the samples. This has already been

reported by Circone et al. (1991) who suggested that it should result from some K-

deficient clay-like layers in the phlogopites. This signal was only observed as a small

shoulder but in some rare cases it showed up to 10% relative signal intensity. It has

only been observed for low Al-contents (x ≤ 0.6), but no correlation between its

occurrence and the fluorine content of the samples has been found. It could be due

to clinochlore which was identified in the XRD patterns of some samples. In any

case, this additional signal could be easily distinguished from those resulting from

phlogopite and did not disturb during analysis of the data.

Figure 4.2 shows a comparison of 29Si MAS NMR spectra of phlogopites with

different F- and Al-contents. The spectra of phlogopites of different F-contents but of

same nominal Al-content xnom do not differ much. However, a change of signal

intensities for constant F-content but increasing Al-content can clearly be seen. At

low Al-contents the Si-Si2Al-signal shows the highest signal intensity, and the ratio of

Si-SiAl2 : Si-Si2Al : Si-Si3 is about 25:50:25. With increasing Al-content, the intensity

of the Si-SiAl2-signal increases until this signal becomes equal to or slightly higher

than the Si-Si2Al signal. At the same time, a fourth signal corresponding to the Si-Al3

environment rises in intensity. All observations indicate that no kind of tetrahedral

ordering is present in these structures except for a short-range ordering controlled by

avoidance of Al-O-Al linkages.

4.2. Ordering of cations in the tetrahedral sheets of phlogopite

65

x =

0.2

x =

0.4

x =

0.6

x =

0.8

y =

1.0

y =

1.2

y =

1.4

y =

1.6

y =

1.8

x =

0.0

7es

tx

= 0

.11

est

x =

0.1

4e

stx

= 0

.03

est

x =

0.1

4e

st

x =

0.1

4e

stx

= 0

.13

est

x =

0.1

0e

stx

= 0

.17

est

x =

0.4

1es

tx

= 0

.23

est

x =

0.2

9e

stx

= 0

.28

est

x =

0.4

3e

st

xt =

0.5

1es

x =

0.4

6e

stx

= 0

.35

est

-80

-10

0-6

0pp

m-8

0-1

00

-60

ppm

-80

-100

-60

ppm

-80

-10

0-6

0pp

m-8

0-1

00-6

0p

pm

-80

-10

0-6

0pp

m-8

0-1

00-6

0p

pm-8

0-1

00

-60

ppm

-80

-10

0-6

0p

pm

-80

-10

0-6

0pp

m-8

0-1

00

-60

ppm

-80

-100

-60

ppm

-80

-10

0-6

0pp

m-8

0-1

00-6

0p

pm

-80

-10

0-6

0pp

m-8

0-1

00

-60

ppm

-80

-10

0-6

0pp

m

nominal x

Figure 4.2. Comparison of 29Si MAS NMR spectra of phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y with

different Al- and F-contents. Below the spectra, the Al-content of the tetrahedral sheets calculated from the

relative signal intensities, xest, is given.

4. Results and discussion

66

From the relative signal intensities the Si/[4]Al-ratio can be calculated using the

following equation (Lipsicas et al., 1984; Sanz and Serratosa 1984):

3

0=n

3

3

0=n

3

[4]

)(3

)( =

Al

Si

nAlQIn

nAlQI

n

n

(4.2)

where Q3 denotes a tetrahedral site linked to three other tetrahedra and In is the

relative signal intensity of a Q3-Si-signal with n Al-atoms as next-nearest-neighbour.

Figure 4.3a shows a plot of this ratio against the nominal Al-content of the gel

composition. The solid black curve indicates the phlogopite Al-content if all of the

starting material had reacted to phlogopite. However, the Si/[4]Al-ratio is always

higher than predicted due to the formation of Al-rich impurity phases like corundum.

If we compare these values to the samples synthesised at 800 °C and reported

by Fechtelkord et al. (2003a) as shown in Figure 4.3a, we see that the data of the

samples synthesised at 600 °C scatter much more. Moreover, at 800 °C synthesis

temperature, clear trends have been observed for the Si/Al-ratio with different

F-contents, which are not visible in the new data. Also, both data sets are in the

same range, indicating that the amount of Al incorporated into the structure is roughly

the same.

The fact that not all Al of the initial gel composition is incorporated into the

phlogopite structure becomes even more visible, if the estimated Al-content

)1(

)3( =x

]4[

]4[

est

Al

SiAl

Si

(4.3)

of the tetrahedral sheet is calculated from the experimental Si/[4]Al-ratio and then

compared to the nominal Al-content of the initial gel, xnom (see Figure 4.3b). For all

samples the experimental value is lower than the nominal x, and this effect becomes

more pronounced with higher initial Al-contents. This estimated Al-content xest is also

given below the corresponding 29Si MAS NMR spectra in Figure 4.2. As could

already be seen from the change in signal intensities, these values do not change

much with different F-content for a given Al-content, but increase as the Al-content of

4.2. Ordering of cations in the tetrahedral sheets of phlogopite

67

uncertainty:(this study)

uncertainty:

Fechtelkord et al. (2003a)

a)

b)

Figure 4.3. a) The Si/[4]Al-ratio calculated from the relative signal intensities in the 29Si MAS NMR spectra plotted

against the nominal Al-content of the initial oxide mixture. The solid black curve indicates the phlogopite

composition if all starting material had reacted to phlogopite. b) Plot of the experimentally derived (additional) Al-

content of the tetrahedral sheets of the phlogopites against the Al-content of the initial gel mixture. The black line

indicates a complete reaction of the starting material to phlogopite.

the nominal gel composition increases. The highest Al-content of the initial gel

compositions was xnom = 0.8, however, in these samples the highest estimated Al-

content xest was far lower with xest ~ 0.45.

4. Results and discussion

68

4.2.2. Hydroxyl-phlogopites and Al- and OH-rich phlogopites

(0.8 ≤ x ≤ 1.6; 1.6 ≤ y ≤2.0)

A number of pure hydroxyl-phlogopite samples have been prepared to allow for a

comparison of F-free and F-containing compositions. Moreover, to investigate the

maximum amount of Al-incorporation into the phlogopite structure at synthesis

conditions of 600 °C and 2 kbar, a few samples with high nominal Al-content

(xnom = 1.0, 1.2, and 1.6) have been prepared. These only contain low amounts of

fluorine (y = 1.6, 1.8, and 2.0) as it can be expected that more F-rich compositions

will not be able to incorporate more Al anyway.

Theoretically, the limit for an incorporation of Al into the phlogopite structure is

x = 1.0 because then a ratio of Si/[4]Al of 1:1 is reached. Incorporating further Al into

the tetrahedral sheets would force Al-atoms to occupy neighbouring tetrahedra which

is expected to be highly energetically unfavourable. However, values of xnom higher

than 1.0 have been chosen for the preparation of gels because not all of the Al

present in the initial gel composition is incorporated into the phlogopite structure, as

has been discussed in Chapter 4.2.1 for samples of lower Al-contents.

The 29Si MAS NMR spectra of the Al-rich phlogopites shown in Figure 4.4 exhibit

the same signals as have been observed for low Al-phlogopites: The Si-Al3, the

Si-SiAl2, the Si-Si2Al, and the Si-Si3 signal at -80, -83, -87, and -90 ppm, respectively.

At Al-contents lower than xest = 0.6 the Si-SiAl2 and the Si-Si2Al signals dominate, but

at higher Al-contents the Si-Al3 signal shows the highest signal intensity. This

indicates that long-range ordering is established, with Si and Al occupying tetrahedral

sites alternately. This kind of ordering also leads to a narrowing of the Si-SiAl3 signal

from 2.7 ppm to 1.3 ppm because the Si environment becomes more uniform

throughout the structure (see Table 6.2 in the Appendix). These results are in good

agreement with the findings of Circone et al. (1991) who reported the same for

hydroxyl-phlogopites coming close to Si/[4]Al-ratios of 1:1.

Still, a higher Al-content of the initial gel composition also means a higher

incorporation of this element into the phlogopite structure. However, a saturation

effect could be observed: For y = 1.8, a maximum value of xest = 0.71 has been

achieved for xnom = 1.2. For higher xnom of 1.6, the value is nearly the same with

xest = 0.68.

4.2. Ordering of cations in the tetrahedral sheets of phlogopite

69

x no

m =

1.0

x nom

= 1

.2x n

om =

1.6

-60

-80

-100

ppm

-60

-80

-100

ppm

y =

2.0

y =

1.6

-60

-80

-100

ppm

-60

-80

-100

ppm

-60

-80

-100

ppm

x nom

= 0

.8

-60

-80

-100

ppm

-60

-80

-100

ppm

x =

0.8

3e

stx

= 0

.68

est

x =

0.8

2es

tx

= 0

.68

est

x =

0.4

2es

tx

= 0

.53

est

x =

0.6

6es

t

y =

1.8

x =

0.7

1es

tx

= 0

.68

est

-60

-80

-100

ppm

-60

-80

-100

ppm

Figure 4.4. Comparison of 29Si MAS NMR spectra of phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y with high

Al- and low F-contents. Below the spectra, the Al-content of the tetrahedral sheets calculated from the relative

signal intensities, xest, is given.

4. Results and discussion

70

uncertainty:(this study)

uncertainty:

Figure 4.5. a) The Si/[4]Al-ratio calculated from the relative signal intensities in the 29Si MAS NMR spectra plotted

against the nominal Al-content of the initial oxide mixture. The solid black curve indicates the phlogopite

composition if all starting material had reacted to phlogopite. b) Plot of the experimentally derived (additional) Al-

content of tetrahedral sheets of the phlogopites against the Al-content of the initial gel mixture. The black line

indicates a complete reaction of the starting material to phlogopite.

4.2. Ordering of cations in the tetrahedral sheets of phlogopite

71

y =

2.0

y =

1.6

y =

1.0

x =

0.8

x =

1.2

-50

-70

-90

-11

0(p

pm)

-50

-70

-90

-11

0(p

pm)

-50

-70

-90

-11

0(p

pm)

x =

0.4

-50

-70

-90

-110

(ppm

)

y =

0.5

x=

0.2

5es

t

x=

0.6

8es

t

x=

0.8

2es

t

x=

0.1

2es

t

x=

0.4

2es

t

x=

0.6

6es

t

x=

0.2

4es

t

x=

0.5

0es

t

x=

0.1

0es

t

x=

0.4

2es

t

Figure 4.6. Comparison of 29Si MAS NMR spectra of phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y with

different Al- and F-contents.

4. Results and discussion

72

Higher Al-contents up to xest = 0.83 have been reached using F-free starting

materials (y = 2.0). In this case the incorporation of Al is closer to the stoichiometric

composition, especially at higher Al-contents (Figure 4.5). The maximum value has

been reached for a nominal xnom of 1.0 and 1.2. For xnom = 1.6, the estimated

Al-content is lowered again, indicating a preferred formation of other phases than

phlogopite at such high Al-contents of the F-free oxide mixture.

The maximum Al-contents found in this study are still lower than the highest value

reached so far, xest = 0.92, reported by Circone et al. (1991). However, these authors

had slightly lower synthesis temperatures of 400-600 °C and higher pressures of 5

kbar.

If one compares the 29Si MAS NMR spectra of F-free phlogopites with those of

phlogopites of different F-contents, as shown in Figure 4.6, it becomes obvious that

the spectra and the amount of Al incorporated into the structure do not change much

with increasing F-content. In contrast, the difference between pure OH-phlogopites

(y = 2.0) and those containing only little F (y = 1.8) is much larger than the difference

between phlogopites of varying F-contents. Therefore, it can be concluded that the

mere presence of F in the mixture is much more important than the exact ratio of

OH/F. As soon as the system contains F, the ability of the phlogopites to incorporate

Al is lowered drastically.

4.2.3. Samples of high F-contents (y < 1.0)

In addition to the samples discussed before, a number of phlogopites with higher

F-contents, i.e., y = 0.8 and y = 0.5 have been prepared. Moreover, two very F-rich

samples with compositions xnom = 0.1, y = 0.2 and xnom = 0.8, y = 0.0, respectively,

have been synthesised and analysed.

The 29Si NMR spectra of these F-rich samples are shown in Figure 4.7. Many of

them have a lower signal to noise ratio than that of the more OH-rich samples

indicating a lower amount of phlogopite formed in these samples. Even for Mg- and

Si-rich compositions the crystallisation of phlogopite seems to be less favourable

under these conditions.

The signals present are the same as have been observed for other compositions,

with the Si-Si3-signal at -90 ppm, the Si-Si2Al-signal at -87 ppm, the Si-SiAl2-signal at

4.2. Ordering of cations in the tetrahedral sheets of phlogopite

73

y =

0.5

-60

-80

-100

ppm

-60

-80

-100

ppm

-60

-80

-100

ppm

x e

st =

0.0

0x

est

= 0

.29

x es

t = 0

.33

x es

t =

0.0

1x

est

= 0

.24

-60

-80

-100

ppm

-60

-80

-100

ppm

-60

-80

-100

ppm

-60

-80

-100

ppm

-60

-80

-100

ppm

-60

-80

-100

ppm

y =

0.8

x =

0.2

x =

0.4

x =

0.6

x =

0.8

x e

st =

0.4

0x

est

= 0

.41

x es

t =

0.3

7x

est

= 0

.11

-60

-80

-100

ppm

x es

t =

0.4

0

nom

inal

x

x =

0.1

y =

0.2

x =

0.8

y =

0.0

Figure 4.7. Comparison of 29Si MAS NMR spectra of F-rich phlogopites with nominal composition

K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y.

4. Results and discussion

74

-83 ppm, and the Si-Al3-signal at -80 ppm, respectively. At low initial Al-contents, no

additional Al has been incorporated into the structure, and the spectra show the

typical Si-Si3:Si-Si2Al:Si-SiAl2 = 25:50:25 distribution. Only for y = 0.8 the amount of

Al incorporated into the tetrahedral sheet is higher with xest = 0.1. As expected, the

amount of Al in the tetrahedral sheets increases at higher Al-contents of the initial

oxide mixture, and the Si environments with one or two Al- atoms dominate the

spectrum. With xest = 0.4 for xnom = 0.8 and y = 0.5/0.8, the maximum real Al-content

of the phlogopites is much lower than that of the nominal composition. However, this

value is still in the same range as that of samples with higher OH-contents (see

Table 6.2 in the Appendix). Only for the water-free composition of y = 0.0 a sharp

decrease of Al in the structure has been observed, where xest = 0.24 even for a

nominal Al-content of 0.8. The only difference between the samples of y = 0.5 and

y = 0.8 is a faster increase of Al incorporated into the structure for the samples of

y = 0.8.

4.2.4. J-formalism and Monte-Carlo simulations

In the previous chapters, the distribution of Si environments in the tetrahedral

sheets of phlogopites with different F- and Al-contents has been discussed.

Computational techniques based on the J-formalism and Monte-Carlo simulations

have been used to obtain additional information on the ordering of ions in the

tetrahedral sheets and to explain the observed ordering schemes.

So far, this has only been done for F-free compositions with Al-contents ranging

from x = 0.0 to the maximum value of x = 1.0. The latter cannot be obtained

experimentally, but for lower Al-contents a comparison between simulation and

experiment is possible. In the Monte-Carlo simulations, the atom configuration of

lowest energy has been determined. From these, the relative amounts of Si

environments in the structure have been calculated and a ‘theoretical NMR spectrum’

has been constructed.

An overview on the J-parameters within a single tetrahedral sheet is given in

Figure 2.14 in Chapter 2.3. In agreement with Loewenstein’s rule of avoidance of Al-

O-Al linkages in tetrahedral configurations (Loewenstein, 1954) the J1-parameter –

describing the interactions between neighbouring tetrahedral sites – is highly

positive, meaning a strong preference for Al-atoms to have only Si-atoms as next-

4.2. Ordering of cations in the tetrahedral sheets of phlogopite

75

J = 1.07 eV1

J = 0.26 eV2

J = 0.13 eV3J = 0.09 eV4

Figure 4.8. Comparison of the tetrahedral intrasheet J-parameters. The error range of the values is ± 0.05 eV.

Figure 4.9. Configuration of lowest energy for ordering of cations in a single tetrahedral sheet of phlogopite with x

= 1.0 (‘eastonite’ composition, K (Mg2Al) (Al2Si2O10) (OH)2). Al-atoms are shown in red, Si-atoms in yellow. Grey

bars indicate Al-Si-neighbour pairs. Only a part of the supercell is shown. Note the defects characterised by Al-Al

neighbour pairs.

4. Results and discussion

76

b)

a)

Figure 4.10.Configurations of lowest energies for cation ordering in a single tetrahedral sheet of phlogopite with

a) x = 0.5 (composition K (Mg2.5Al0.5) (Al1.5Si2.5O10) (OH)2) and b) x = 0.25 (composition

K (Mg2.25Al07.5) (Al1.75Si2.25O10) (OH)2). Al-atoms are shown in red, Si-atoms in yellow. Grey bars indicate Al-Si-

neighbour pairs. Only a part of the supercell is shown.

4.2. Ordering of cations in the tetrahedral sheets of phlogopite

77

nearest-neighbours. With increasing distance the value of the J-parameters

decreases, until they do not influence the ordering in the sheets anymore (Figure

4.8).

The large influence of the 1J -parameter becomes clearly visible when regarding

only a single tetrahedral sheet of a layer package as shown in Figure 4.9 for x = 1.0.

Every Si-atom is surrounded by three Al-atoms in the neighbouring tetrahedra and

vice versa. Domains are formed separated by an area in which Loewenstein’s rule is

not preserved anymore and two Al-atoms occupy directly neighboured tetrahedra.

With this ordering pattern, only the Si-Al3 signal should appear in the 29Si NMR

spectrum. However, such high Al-contents cannot be obtained experimentally to

verify this observation.

At lower Al-contents, the ordered Si-Al3 domains with a ratio of Si/[4]Al = 1:1 are

still present, although the composition has changed (Figure 4.10). In order to

preserve the overall composition, areas of phlogopite composition in the narrower

sense (Si/[4]Al = 3:1) appear, more and more replacing the Si-Al3 domains with

decreasing Al-content. Partly, a new type of ordering is established in these Al-poor

areas, with Al-atoms occupying sites which are linked by the 3J -interaction, i.e. on

opposite sites of the hexagonal rings (upper right part of Figure 4.10b). For x = 0.0,

only areas with ordering on sites linked by 3J -interaction remain and long-range

ordering is established again (Figure 4.11).

The ordering on 3J -coupled sites is forced by composition: The 4J -parameter

has the lowest value of all the interactions active within one tetrahedral sheet.

However, placing Al-atoms on these positions leads to a distribution of only one Al-

atom per hexagonal ring. This is not possible because the lowest value of Si/[4]Al is

3:1, so that Al-atoms would necessarily also have to occupy sites coupled by the 2J -

interaction. It is therefore more favourable for Al to order on the 3J -coupled sites on

opposite sides of the hexagonal rings. As has already been observed for the Si-Al3

ordering at high Al-contents, domains occur between which only one Al-atom per

hexagonal ring can be found.

If we compare these results with the experimental 29Si MAS NMR spectra as

shown in Figure 4.12 and Table 4.2, it becomes clear that experiment and theory

agree well at high Al-contents but diverge for low-Al compositions. At x = 0.82, the

4. Results and discussion

78

Figure 4.11. Configurations of lowest energy for cation ordering in a single tetrahedral sheet of phlogopite with

x = 0.0 (composition K Mg3 (AlSi3O10) (OH)2). Al-atoms are shown in red, Si-atoms in yellow. Every Al-atom has

three Si-atoms as next-nearest-neighbours, while every Si-atom is surrounded by two Si-atoms and one Al-atom

in the neighbouring tetrahedra. a) Grey lines indicate Si-Al neighbour pairs. b) The J3-interactions connecting Al-

atoms are marked by grey lines.

4.2. Ordering of cations in the tetrahedral sheets of phlogopite

79

spectrum shows a high intensity for the Si-Al3 signal and only very low intensities for

the signals of Si environments with lesser numbers of Al-neighbours. This can be

well described by the Si-Al3 domains making up the largest part of the structure, while

Si-SiAl2, Si-Si2Al and Si-Si3 environments are only present at the borders between

different domains and in the smaller disordered parts of the structure.

As the Al-content is lowered to x = 0.67, the Si-Al3 signal decreases because the

corresponding ordered parts of the structure become smaller. Meanwhile, the other

three signals increase in intensity as the areas of low Al-contents replace the ordered

domains. However, the agreement is not as good any more: The relative amount of

Si-SiAl2 environments is highly underestimated in the simulations. These are mostly

found at the borders between the ordered and the disordered domains which could

mean that the cluster size is still too large in the simulations. Reducing the cluster

size would mean an increase in border area and thus an increase in the Si-SiAl2

signal intensity.

The same problems seem to occur for x = 0.25, where the relative amount of the

Si-Si2Al environments is underestimated in the simulations. For a composition of

x = 0.0, the simulations are not able to describe the ordering observed in the 29Si

MAS NMR spectra. According to the Monte-Carlo simulations, Al should order on the

3J -connected sites, as has been discussed above, leading to long-range ordering

again. This type of ordering has been reported before by Palin and Dove (2004) for

tetrahedral sheets of the dioctahedral equivalent to phlogopite, the mica muscovite (K

Al2 (AlSi3O10) (OH)2).

All the Al-atoms are now surrounded by Si in the neighbouring tetrahedra only,

whereas Si-atoms have two Si-atoms and one Al-atom as next-nearest-neighbours.

In the corresponding 29Si MAS NMR spectra, only the Si-Si2Al signal should be

Table 4.2. Comparison of relative numbers of Si environments in the tetrahedral sheets of phlogopite.

x simulations experiment

Si (3Al) Si (2Al) Si (1Al) Si (0Al) Si (3Al) Si (2Al) Si (1Al) Si (0Al)

0.25 10 43 21 27 7 34 43 13

0.68 56 19 13 13 44 34 17 5

0.82 73 12 8 7 61 29 9 0

4. Results and discussion

80

-70 -80 -90 -100 (ppm) -80-70 -90 -100 (ppm)

-70 -80 -90 -100 (ppm) -70 -80 -90 -100 (ppm)

-70 -80 -90 -100 (ppm)

x = 0.25

x = 0.68

x = 0.82

-70 -80 -90 -100 (ppm) -70 -80 -90 -100 (ppm)

x = 0.00

Figure 4.12. Comparison of experimental 29Si MAS NMR spectra (right) and theoretical ones derived from the

Monte-Carlo simulation results (left) for phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)2. For x = 0.0, the 29Si MAS

NMR spectrum of a F-bearing phlogopite is shown as no F-free samples have been available for analysis.

4.2. Ordering of cations in the tetrahedral sheets of phlogopite

81

visible, having a relative signal intensity of 98 %. Two small shoulders result from the

Si-Si3 and the Si-SiAl2 environments present at the borders between the domains,

and both have a signal intensity of 1 %.

The experimentally observed 29Si MAS NMR spectrum, however, differs

completely from the theoretical one. The relative signal intensities of the Si-Si3 and

the Si-SiAl2 environments are much higher, leading to a ratio of

Si-SiAl2 : Si-Si2Al : Si-Si3 of 25:50:25 instead of 1:98:1. This means that long-range

ordering is not present, but Al is distributed statistically in pure Mg-phlogopite,

regarding one tetrahedral sheet only.

One possible explanation for the deviation between simulations and experiment

could be that pure Mg-phlogopites have only been synthesised for F-bearing

compositions, whereas fluorine is not considered in the simulations. However, this is

not likely as the 25:50:25 signal distribution has been observed in all low-Al samples

independently from the exact F-content. These findings are also in agreement with

the results of Circone et al. (1991) who proposed that only short-range ordering due

to the avoidance of Al-O-Al linkages should be present in phlogopites of low Al-

contents.

4. Results and discussion

82

4.3. Ordering of cations and anions in the octahedral sheets of

phlogopite

4.3.1. Samples of intermediate to low F-contents (1.0 ≤ y ≤ 1.8)

Both 1H and 19F MAS NMR spectroscopy give the opportunity to probe the OH/F

environment in the octahedral sheet and to obtain information on the distribution of

Mg/Al in this sheet. Theoretically, several signals corresponding to different amounts

of Al in the OH/F coordination sphere may occur in both types of spectra: OH/F-Mg3,

OH/F-Mg2Al, OH/F-MgAl2, and OH/F-Al3.

4.3.1.1. 1H MAS NMR spectroscopy

In the 1H MAS NMR spectra shown in Figure 4.13, only two of the expected

signals have been observed: the H-OMg3 signal at about 0.5 ppm and the H-OMg2Al

at about 1.8 ppm. If the overall Al-content of the samples is low, the amount of Al-

atoms in the octahedral sheet is not high enough for H-OMgAl2 and H-OAl3 signals to

appear, but with increasing Al-content it should become more and more likely for two

aluminium atoms to occupy neighbouring octahedra. Those signals were always

absent, regardless of the Al-content. This implies an avoidance of Al in neighbouring

octahedra, similar to the avoidance of Al-O-Al linkages in the tetrahedral layers.

The relative intensity of the H-OMg3 and the H-OMg2Al signals clearly changes

with composition. With more Al being incorporated into the structure (at constant F-

content), the H-OMg2Al signal intensity clearly increases, becoming even higher than

that of the H-OMg3 signal. For F-rich compositions the H-OMg2Al signal intensity is

higher than for OH-rich samples (at constant Al-content), although one might expect

the opposite as OH-rich phlogopites are able to incorporate larger amounts of Al.

Fechtelkord et al. (2003a) reasoned that the strong preference of OH to be co-

ordinated by Al instead of Mg should be responsible for this observation. At

intermediate OH contents, Al is gathered around the OH-groups instead of being

distributed statistically. For OH-rich compositions the Al-content is not high enough

for a co-ordination of all hydroxyl groups by Mg2Al. Therefore, a larger number of OH

is forced to occupy Mg3-sites, too.

4.3. Ordering of cations and anions in the octahedral sheets of phlogopite

83

x =

0.2

x =

0.4

x =

0.6

x =

0.8

y =

1.0

y =

1.2

y =

1.4

y =

1.6

y =

1.8

Al/(

Mg+

Al)

= 0

.49

Al/(

Mg+

Al)

= 0

.38

Al/(

Mg+

Al)

= 0

.47

Al/(

Mg

+A

l) =

0.5

3A

l/(M

g+A

l) =

0.5

5

Al/(

Mg+

Al)

= 0

.54

Al/(

Mg+

Al)

= 0

.46

Al/(

Mg+

Al)

= 0

.44

Al/(

Mg+

Al)

= 0

.31

Al/(

Mg+

Al)

= 0

.34

Al/(

Mg+

Al)

= 0

.42

Al/(

Mg+

Al)

= 0

.44

Al/(

Mg+

Al)

= 0

.54

Al/(

Mg+

Al)

= 0

.43

Al/(

Mg+

Al)

= 0

.41

Al/(

Mg+

Al)

= 0

.33

Figure 4.13. 1H MAS NMR spectra of phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y with different OH- and

Al-contents. Below the spectra the ratio I[H-OMg2Al]/(I[H-OMg2Al] + I[H-OMg3]) is given, abbreviated as

‘Al/(Mg+Al)’.

4. Results and discussion

84

uncertainty:(this study)

OH

Al/(

OH

Al+

OH

Mg

)

uncertainty:

OH

Al/(

OH

Al+

OH

Mg

)

a)

b)

Figure 4.14. Plot of the relative signal intensity of the H-OMg2Al signal against the Al-content x estimated from 29Si MAS NMR spectra. The solid line represents a statistical distribution of ions in the octahedral sheet. a) Data

of this study only, b) comparison of the 600 °C data (black symbols) to the 800 °C data of Fechtelkord et al.

(2003a, grey symbols).

4.3. Ordering of cations and anions in the octahedral sheets of phlogopite

85

This non-statistical distribution of Mg and Al in the octahedral sheet is more

obvious in plots of the relative signal intensity of the H-OMg2Al signal,

I[H-OMg2Al]/(I[H-OMg2Al] + I[H-OMg3]), against the Al-content of the phlogopites

estimated from the 29Si MAS NMR spectra, xest (Figure 4.14a, see also data given in

Table 6.3 in the Appendix). Assuming that Al is incorporated into both sheets equally

(according to Tschermak’s substitution), xest should also be the Al-content of the

octahedral sheet. H-OMgAl2 and H-OAl3-signals have not been observed which

means only H-OMg3 and H-OMg2Al environments are present. Two of the sites co-

ordinating each hydroxyl-group are always occupied by Mg, and Mg ↔ Al exchange

only takes place on the third site. For compositions of xest = 0.0, this site is occupied

by Mg only, while at xest = 1.0 all sites are occupied by Al. Therefore, the H-OMg2Al

and the H-OMg3 signal intensities directly correlate to xest and (1-xest), respectively,

and the Al-content of the octahedral sheet is equal to the relative H-OMg2Al signal

intensity:

estestest

est xxx

x

OMgHIAlOMgHI

AlOMgHI

)1(]) [][(

] [

32

2 (4.4)

This relationship is valid for a statistical distribution of ions in the octahedral sheet.

However, Figure 4.14a shows that the H-OMg2Al signal intensity is much higher than

expected, confirming the non-statistical Mg/Al distribution with Al preferring OH

environments over F environments. At high Al-contents of xest > 0.40 the

experimental data approach the theoretical values and the ordering pattern observed

before is not present any more.

Fechtelkord et al. (2003a) published similar data for their phlogopite samples

synthesised at 800 °C (1073 K). However, in their publication these values were

plotted against the nominal Al-content of the sample and not the real Al-content of

the phlogopites. The amount of Al incorporated into these phlogopites has been

calculated from their data according to equation (4.3) in order to compare the

resulting curves to the ones obtained in this study (Figure 4.14b). The data are

roughly in the same range for x > 0.30, but at lower Al-contents there is a larger

discrepancy between both data sets. The low-T phlogopites already show a strong

enrichment of Al-atoms around OH-groups for very low overall aluminium contents,

while the samples synthesised at 800 °C show this only for OH-rich compositions of

4. Results and discussion

86

y = 1.8. A possible explanation for this observation could be a higher degree of

ordering at lower temperatures.

4.3.1.2. 19F MAS NMR

As fluorine substitutes for OH in the phlogopite structure we expect the same two

signals as have already been observed for 1H: The first one is due to a co-ordination

by three Mg-ions, and the other one results from an environment with two Mg and

one Al. According to Huve et al. (1992a,b) a signal observed at -175 ppm can be

assigned to the F-Mg3 environment in the phlogopite structure, while another signal

located at -150 ppm results from a co-ordination by 2Mg and 1 Al (Figure 4.15).

For the compositions discussed in this chapter hardly any signals resulting from

impurity phases have been observed. A sharp signal has been found at -177 to

-179 ppm for xnom = 0.2 / y = 1.2 and xnom = 0.6 / y = 1.8. It is not clear where this

signal results from, and its appearance does not depend on the sample composition.

Another signal with a broad Lorentzian lineshape has been found in the spectra of

intermediate F-contents (y = 1.0) and high Al-contents (xnom = 0.6 and 0.8). It makes

up a large amount of the whole signal intensity with 34 and 45 % relative signal

intensity, respectively, and possibly results from a F- and Al-containing amorphous

part of the sample. However, the intensity of this signal might be overestimated due

to short spin-lattice relaxation of the corresponding compound, so that the real

amount of amorphous material is lower.

The signal intensity of spinning sidebands resulting from the two phlogopite

signals, marked by asterisks in the spectra, is very low, indicating low chemical shift

anisotropy due to the nearly axial symmetry of the F-Mg3 environment. The intensity

of the F-Mg2Al signal has been very low, so the corresponding spinning sidebands of

this signal could not be observed.

Similar to the 1H MAS NMR spectra the relative intensities of the F-Mg3 and the

F-Mg2Al signals change with composition (Figure 4.15, Table 6.4 in the Appendix). At

low Al contents hardly any F-Mg2Al environments are present in the structure

because F strongly prefers a co-ordination by three Mg-ions. As the amount of Al in

the structure increases, more and more F-ions are forced to be co-ordinated by Al,

too, leading to an increase of the corresponding signal intensity. However, the

amount of F-Mg2Al environments is still very low compared to that of the H-Mg2Al

environments due to the strong F-Al-avoidance previously observed in micas. This is

4.3. Ordering of cations and anions in the octahedral sheets of phlogopite

87

x =

0.2

x =

0.4

x =

0.6

x =

0.8

y =

1.0

y =

1.2

y =

1.4

y =

1.6

y =

1.8

nominal x

-120

-160

-20

0pp

m

-120

-160

-200

ppm

-120

-16

0-2

00p

pm

-120

-160

-200

ppm

-120

-160

-200

ppm

-120

-160

-200

ppm

-120

-160

-200

ppm

-120

-160

-200

ppm

-120

-160

-200

ppm

-12

0-1

60-2

00pp

m

-120

-160

-200

ppm

-12

0-1

60-2

00pp

m

-120

-160

-200

ppm

-120

-16

0-2

00p

pm

Al/(

Mg+

Al)

= 0

.08

Al/(

Mg+

Al)

= 0

.11

Al/(

Mg+

Al)

= 0

.04

Al/(

Mg+

Al)

= 0

.22

Al/(

Mg+

Al)

= 0

.13

Al/(

Mg+

Al)

= 0

.07

Al/(

Mg+

Al)

= 0

.04

Al/(

Mg+

Al)

= 0

.01

Al/(

Mg+

Al)

= 0

.04

Al/(

Mg+

Al)

= 0

.05

Al/(

Mg+

Al)

= 0

.03

Al/(

Mg+

Al)

= 0

.02

Al/(

Mg+

Al)

= 0

.05

Al/(

Mg+

Al)

= 0

.05

Al/(

Mg+

Al)

= 0

.06

Al/(

Mg+

Al)

= 0

.12

Al/(

Mg+

Al)

= 0

.07

-12

0-1

60-2

00pp

m

-120

-160

-200

ppm

-120

-160

-200

ppm

**

**

**

**

**

**

**

** *

**

***

**

** *

* **

***

*

**

**

Figure 4.15. Comparison of 19F MAS NMR spectra of phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y with

different Al- and F-contents. Below the spectra the ratio I[F-Mg2Al]/(I[F-Mg2Al] + I[F-Mg3]) is given, abbreviated as

‘Al/(Mg+Al)’. Spinning sidebands are marked by asterisks.

4. Results and discussion

88

F Al/(

F Al+

FM

g)

uncertainty:

Figure 4.16. Plot of the relative intensity of the F-OMg2Al signal against the Al-content x estimated from 29Si MAS

NMR spectra. The solid line represents a statistical distribution of ions in the octahedral sheet.

also an explanation for the slight decrease of the F-Mg2Al signal with increasing OH-

content. At low F-contents, the amount of Mg3-sites is high enough to avoid F-Mg2Al

co-ordination, but with increasing F-content, F-ions are forced to occupy Mg2Al sites,

too.

At high OH-contents, i.e., y = 1.6 and y = 1.8, the signal to noise ratio becomes

significantly worse with increasing Al-content of the initial gel composition. At such

high Al-contents in combination with only low amounts of F available in the oxide

mixture, the incorporation of F into the phlogopite structure is very low due to the

Al-F-avoidance mentioned above.

In Chapter 4.3.1.1 it has been shown that the Al-content of the hydroxyl

environment is given directly by the relative H-OMg2Al signal intensity. Analogous the

following equation should prove true for a statistical distribution of Al and Mg in the

octahedral sheets:

4.3. Ordering of cations and anions in the octahedral sheets of phlogopite

89

estxMgFIAlMgFI

AlMgFI

])[] [(

] [

3 2

2. (4.5)

This relationship is maintained at low values of xest, but with increasing Al-content the

experimental data strongly deviate from the theoretical values (Figure 4.16).

Especially at high overall Al-contents the F environment contains less Al then

expected indicating a non-statistical distribution of ions in the octahedral sheet. This

confirms the results of Papin et al. (1997) and Fechtelkord et al. (2003a) who

suggested a preference for F to be co-ordinated by Mg only.

It could be expected that larger Al-contents might make it more difficult for F to be

co-ordinated by Mg only, and the deviation from a statistical distribution should be

less pronounced. However, this is not the case. One possible explanation might be a

rather low amount of F-incorporation into these Al-rich phlogopites due to excess

water during the synthesis and F-Al-avoidance. In this case the amount of Mg-atoms

in the octahedral sheet was still sufficient to ensure a co-ordination of F mostly by

three Mg. Indeed, the F-contents of high-Al phlogopites have found to be relatively

low (for details see Chapter 4.3.2.2).

4.3.2. Hydroxyl-phlogopites and Al- and OH-rich phlogopites

(0.8 ≤ x ≤ 1.6; 1.6 ≤ y ≤2.0)

4.3.2.1. 1H MAS NMR spectroscopy

The 1H MAS NMR spectra of OH- and Al-rich phlogopites shown in Figure 4.17

are in principle similar to those discussed in Chapter 4.3.1.1 for samples of lower Al-

and OH-contents. Again, parameters obtained from fits of the spectra are given in

Table 6.3 in the Appendix. They show the same rise of relative H-OMg2Al signal

intensity with increasing Al-content of the initial gel composition. A slight decrease

has only been observed for y = 2.0 where the sample of xnom = 1.6 shows a lower

ratio of I[H-OMg2Al]/(I[H-OMg2Al] + I[H-OMg3]) than the sample of xnom = 1.2. This

agrees with the 29Si MAS NMR spectroscopic results discussed in Chapter 4.2.2

where a slight decrease of the amount of Al incorporated into the tetrahedral sheet

has been found for this composition.

4. Results and discussion

90

y =

1.6

y =

1.8

y =

2.0

x =

0.8

x =

1.0

x =

1.2

x =

1.6

Al/(

Mg+

Al)

= 0

.66

Al/(

Mg+

Al)

= 0

.55

Al/(

Mg+

Al)

= 0

.65

Al/(

Mg+

Al)

= 0

.78

Al/(

Mg+

Al)

= 0

.74

Al/(

Mg+

Al)

= 0

.81

Al/(

Mg+

Al)

= 0

.71

Al/(

Mg+

Al)

= 0

.54

Al/(

Mg+

Al)

= 0

.80

Al/(

Mg+

Al)

= 0

.68

Figure 4.17. Comparison of 1H MAS NMR spectra of OH- and Al-rich phlogopites

K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y. Below the spectra the ratio I[H-OMg2Al]/(I[H-OMg2Al] + I[H-OMg3]) is given,

abbreviated as ‘Al/(Mg+Al)’.

4.3. Ordering of cations and anions in the octahedral sheets of phlogopite

91

In Chapter 4.3.1.1 it has been shown that in case of a statistical distribution of

Mg/Al and OH/F in the octahedral sheets, the relative H-OMg2Al signal intensity

should be equal to the additional Al-content of the tetrahedral sheet, xest. A non-

statistical distribution of ions was observed for compositions of low Al- (xnom < 0.8)

and intermediate to low F-contents (1.0 ≤ y ≤1.8). With increasing amounts of Al in

the structure the deviation from a statistical distribution was found to be less

pronounced. The same behaviour is true for the OH- and Al-rich samples discussed

in this chapter. In Figure 4.18 a plot of the relative H-OMg2Al signal intensity against

xest is shown. For Al-rich compositions of xest ≥ 0.7 the experimental data are not

different from the theoretical values any more.

The F-free samples of y = 2.0 have to be considered separately. As these

phlogopites do not contain any F in the octahedral sheets, no ordering of Mg/Al

between OH- and F environments is possible, and the relative H-OMg2Al signal

intensity is equal to the overall Al-content of the octahedral sheet. This enables a

direct comparison of the amount of [6]Al and additional [4]Al: If all Al was incorporated

via Tschermak’s substitution ([6]Mg[4]Si ↔ [6]Al[4]Al), all y = 2.0 data points should be

located on the straight line in Figure 4.18. This is only the case for Al-rich

compositions, at low Al-contents, the amount of [6]Al is higher than that of [4]Al. This

indicates that other mechanisms of Al-incorporation different from Tschermak’s

substitution might take place at these compositions. However, it is also possible that

Loewenstein’s rule is not valid at low Al-contents, and the amount of [4]Al is

underestimated due to Al-O-Al linkages.

There is an indication for a slight shift of both the H-OMg3 and the H-OMg2Al

signal position to more-positive values at very high Al-contents (Figure 4.19.). This

effect might be within the error range of the signal position determination. However,

this shift was also observed in the 19F MAS NMR spectra, where it can be seen more

clearly due to the larger chemical shift range covered in these spectra. As F and

hydroxyl-groups occupy the same crystallographic site it can be assumed that the

driving force for the signal position shift is the same for both anions. Possibly, the

signal position shift results from an overlapping of several signals on this spectral

position which cannot be resolved in our spectra (see discussion of 19F MAS NMR

spectra in Chapter 4.3.2.2).

4. Results and discussion

92

OH

Al/(

OH

Al+

OH

Mg

)

uncertainty:

Figure 4.18. Plot of the relative signal intensity of the H-OMg2Al signal against the Al-content xest estimated from 29Si MAS NMR spectra. The solid line represents a statistical distribution of ions in the octahedral sheet.

uncertainty:

a) b)

uncertainty:

Figure 4.19. Plot of the H-OMg2Al (a) and the H-OMg3 (b) signal position as a function of the Al-content of the

estimated Al-content of the phlogopites. Tolerances have been estimated by changing parameters manually

observing χ2 until a distinct change of χ2 took place.

4.3. Ordering of cations and anions in the octahedral sheets of phlogopite

93

4.3.2.2. 19F MAS NMR spectroscopy

The 19F MAS NMR spectra of Al-rich and OH-rich phlogopites are shown in

Figure 4.20, and the relative F-Mg2Al signal intensity is given below the

corresponding spectra. For compositions of y = 1.6 the spectra do not differ much

from those of phlogopites containing less Al shown in Chapter 4.3.1.2. Furthermore,

the relative F-Mg2Al signal intensity remains more or less the same for Al-contents of

xnom = 0.8, 1.0, and 1.2. Again, this value should be in the range of xest, but the

experimental data strongly deviate from the theoretical values (see Table 6.4. in

Appendix for details). The amount of Al co-ordinating F is far lower than expected for

a statistical distribution of ions in the octahedral sheet due to a preference of F to be

surrounded by Mg only.

However, the spectra of even more OH-rich compositions (y = 1.8) exhibit an

additional resonance at -157 ppm not present in the other spectra. In the most Al-rich

sample of xnom = 1.6, this signal makes up about 30% relative signal intensity in the 19F MAS NMR spectrum. So far, a signal at this position has only been observed for

F-rich and Al-rich compositions of y = 0.5 and xnom > 0.5 (Chapter 4.3.3.2), and

Fechtelkord et al. (2003a) reasoned it should result from potassium aluminium

hexafluoride, K3AlF6*0.5H2O.

At first glance these results may seem contradictory, as both types of sample

have a completely different starting composition. However, for both the situation is

similar: Because not all material from the initial gel mixture reacts to phlogopite, the

Al-rich and the F-rich samples contain excess Al and F, respectively. In the latter, the

large amount of F prevents extensive incorporation of Al into the structure, and the

high amounts of F and Al in the residual oxide mixture lead to the formation of

potassium aluminium hexafluoride. In contrast, the samples of y = 1.8 contain only

low amounts of F. However, at very high Al-contents hardly any F is incorporated into

the phlogopite structure, again leading to an enrichment of F and Al in the remaining

gel.

In contrast to the 1H MAS NMR spectra (see Chapter 4.3.2.1.), the signal shift to

more-positive values with increased Al-content is clearly visible in the 19F MAS NMR

spectra. This is due to the larger chemical shift range of several hundred ppms

covered in the 19F MAS NMR spectra compared to only about 10 ppm chemical shift

range for 1H.

4. Results and discussion

94

y =

1.6

y =

1.8

x =

0.8

x =

1.0

x =

1.2

x =

1.6

-12

0-1

60

-20

0pp

m-1

20-1

60

-20

0pp

m-1

20-1

60-2

00

ppm

-120

-160

-200

ppm

-120

-16

0-2

00

ppm

Al/(

Mg+

Al)

= 0

.17

Al/(

Mg+

Al)

= 0

.12

Al/(

Mg+

Al)

= 0

.15

Al/(

Mg+

Al)

= 0

.11

Al/(

Mg+

Al)

= 0

.28

**

**

** *

**

*

Figure 4.20. Comparison of 19F MAS NMR spectra of OH- and Al-rich phlogopites

K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y. Below the spectra the ratio I[F-Mg2Al]/(I[F-Mg2Al] + I[F-Mg3]) is given,

abbreviated as ‘Al/(Mg+Al)’. Spinning sidebands are marked by asterisks.

4.3. Ordering of cations and anions in the octahedral sheets of phlogopite

95

The position of the F-Mg3 signal is shifted from -175 ± 0.4 ppm to -173 ± 0.4 ppm

with increasing Al-content, while the position of the F-Mg2Al signal changes from

-151 ± 0.4 ppm at low Al-contents to -149 ± 0.4 ppm at high Al-contents (Figure 4.21

a,b; Table 6.4 in the Appendix). There is also an indication for an increase of the full

width at half maximum (FWHM) of both signals (Figure 4.21 c,d).

a)

c)

uncertainty:

uncertainty:

d)

uncertainty:

b)uncertainty:

Figure 4.21. Position and full width at half maximum (FWHM) of 19F MAS NMR signals versus estimated Al-

content of phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y. a)+c) F-Mg3 signal. b)+d) F-Mg2Al signal. Tolerances

have been estimated by changing parameters manually observing χ2 until a distinct change of χ2 took place.

Both effects can be explained by small changes in the electron density at the 19F

nuclei due to the different electro-negativities of Mg (1.2) and Al (1.5) (Pauling, 1960).

If Mg is replaced by Al, the electron density at the 19F nucleus decreases and thus,

this nucleus is de-shielded, giving a less-negative signal in the 19F MAS NMR

spectrum. If more and more octahedra are occupied by Al this also influences the

electron density at those OH/F positions which are co-ordinated by Mg only, and thus

a slight shift to the less-negative side of the spectrum may be possible.

4. Results and discussion

96

-120 -140 -160 -180 -200 ppm

* ***

Figure 4.22. 19F MAS NMR spectrum of sample with nominal composition of xnom = 1.2 and y = 1.6 showing a

splitting of the F-Mg3 signal at -175 ppm into two separate signals. Spinning sidebands are marked by asterisks.

However, geometrical reasons must be considered, too. Upon replacement of

Mg2+ by the smaller Al3+ the lateral dimensions of the octahedral sheet decrease. At

the same time tetrahedral ditrigonal rotation is increased to compensate for the

resulting lateral misfit between octahedral and tetrahedral sheets. On the atomic

level, these changes become noticeable already at low Al-contents, as they lead to a

greater variation in bond lengths and bond angles. This in return may also be

responsible for the observed shift and broadening of the 19F resonances. On a larger

scale the Mg ↔ Al substitution leads to a lowering of the a and b lattice parameters

(Chapter 4.7).

Another explanation is the presence of four 19F signals resulting from the F-Mg3

environment, instead of only one. Huve et al. (1992) proposed that four different F

environments should contribute to the signal at about -176 ppm: Two of them should

result from F with one OH-group in the same octahedron, either in cis- or in trans-

position, the other two from octahedrons having two F-atoms either in cis- or in trans-

position. A change in these environments with changing phlogopite composition will

also slightly shift the position of the whole signal made up by the single ones. The

same is in principle true for the Mg2Al signal. Indeed, at very high Al-contents a

splitting of the F-Mg3 signal into two individual signals has been observed for one of

the samples (Figure 4.22).

4.3. Ordering of cations and anions in the octahedral sheets of phlogopite

97

4.3.3. Samples of high F-contents (y < 1.0)

4.3.3.1. 1H MAS NMR spectroscopy

In Figure 4.23 a comparison of 1H MAS NMR spectra of F-rich samples (y = 0.0,

0.2, 0.5, 0.8) to samples of intermediate (y = 1.0) and low (y = 1.6) F-contents is

given. For low to intermediate F-contents it has been shown in the previous chapters

that OH-groups strongly prefer a co-ordination by two Mg and one Al instead of three

Mg. This is also true for the F-rich compositions.

All Al-rich samples show higher H-OMg2Al signal intensities than low-Al samples

because more Al is available to co-ordinate the OH-groups. However, at

compositions of y ≤ 1.0, the ratio of the H-OMg2Al signal intensity to the whole signal

area is approximately the same for xnom = 0.6 and 0.8. This indicates that a saturation

level is reached.

At the same time, the number of H-OMg2Al environments is higher for F-rich

samples than for compositions with less F, although the Al-content of these

phlogopites is not much different. At high OH-contents of y = 1.6, the amount of Al

incorporated into the structure is too low for all OH-groups to have Mg2Al

environments. When going to y = 0.8 and 0.5, the number of hydroxyl-groups is

lowered whereas the Al-content does not change much. Therefore, a higher

percentage of OH is co-ordinated by Mg and Al instead of Mg only.

A sharp decrease in the H-OMg2Al signal intensity is only visible for very high

F-contents of y = 0.2 and 0.0. For these samples a lower F-content has been found in

the 29Si MAS NMR spectra, leading to lower Al-contents in the proton environments,

too.

4.3.3.2. 19F MAS NMR spectroscopy

In Figure 4.24 the 19F MAS NMR spectra of the F-rich phlogopites are shown.

The F-Mg3 signal still dominates the spectrum, however, compared to the spectra of

more OH-rich phlogopites (Figure 4.20) these spectra show a slightly higher relative

intensity of the F-Mg2Al signal. For example, for a constant nominal composition of

xnom = 0.6, I[F-Mg2Al]/(I[F-Mg2Al]+I[F-Mg3]) increases from 0.07 ± 0.05 for y = 1.8 to

0.20 ± 0.05 for y = 0.5.

4. Results and discussion

98

y

= 0

.8y

= 0

.5

x =

0.2

x =

0.4

x =

0.6

x =

0.8

y =

1.0

y =

1.6

Al/(

Mg+

Al)

= 0

.48

Al/(

Mg+

Al)

= 0

.43

Al/(

Mg

+A

l) =

0.2

9

Al/(

Mg

+A

l) =

0.5

0A

l/(M

g+

Al)

= 0

.54

Al/(

Mg

+A

l) =

0.6

3A

l/(M

g+A

l) =

0.7

0

Al/(

Mg

+A

l) =

0.6

7A

l/(M

g+

Al)

= 0

.67

Al/(

Mg

+A

l) =

0.5

4

Al/(

Mg+

Al)

= 0

.49

Al/(

Mg+

Al)

= 0

.53

Al/(

Mg

+A

l) =

0.5

4

Al/(

Mg+

Al)

= 0

.34

Al/(

Mg

+A

l) =

0.4

2

Al/(

Mg+

Al)

= 0

.44

Al/(

Mg+

Al)

= 0

.54

Figure 4.23. Comparison of 1H MAS NMR spectra of phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y. Below the

spectra the ratio I[H-OMg2Al]/(I[H-OMg2Al] + I[H-OMg3]) is given, abbreviated as ‘Al/(Mg+Al)’.

4.3. Ordering of cations and anions in the octahedral sheets of phlogopite

99

y =

0.8

y =

0.5

x =

0.2

x =

0.4

x =

0.6

x =

0.8

-120

-160

-20

0pp

m

-120

-16

0-2

00p

pm-1

20-1

60-2

00pp

m

-120

-160

-200

ppm

-120

-160

-200

ppm

-120

-160

-200

ppm

Al/(

Mg+

Al)

= 0

.09

Al/(

Mg+

Al)

= 0

.06

Al/(

Mg+

Al)

= 0

.02

Al/(

Mg+

Al)

= 0

.15

Al/(

Mg+

Al)

= 0

.10

Al/(

Mg+

Al)

= 0

.18

Al/(

Mg+

Al)

= 0

.16

Al/(

Mg+

Al)

= 0

.20

Al/(

Mg+

Al)

= 0

.07

Al/(

Mg+

Al)

= 0

.18

x =

0.1

y =

0.2

x =

0.8

y =

0.0

-120

-160

-200

ppm

-120

-160

-200

ppm

-120

-160

-200

ppm

-12

0-1

60-2

00pp

m

**

**

**

** *

**

**

**

**

**

* **

**

**

Figure 4.24. Comparison of 19F MAS NMR spectra of F-rich phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y.

Below the spectra the ratio I[F-Mg2Al]/(I[F-Mg2Al] + I[F-Mg3]) is given, abbreviated as ‘Al/(Mg+Al)’. Spinning

sidebands are marked by asterisks.

4. Results and discussion

100

This seems to be in contradiction to the observation that at the same time the

relative H-OMg2Al signal intensity is also higher for F-rich compositions than for OH-

rich phlogopites. However, this can be explained by the preference of Al to co-

ordinate hydroxyl-groups. At high OH-contents, most Al-atoms are found in OH

environments, and F is surrounded mostly by Mg only. Nevertheless, the large

number of OH-groups makes it necessary that H-OMg3 environments are also found

in larger numbers. If the amount of OH-groups decreases, these H-OMg3

environments will disappear and the relative number of H-OMg2Al environments

increases. If the OH-content is sufficiently low, Al-atoms are forced to also co-

ordinate F leading to an increase in relative intensity of the F-Mg2Al signal.

The spectra also show a higher number of signals resulting from impurity phases

than the spectra of phlogopites containing less F. For y = 0.5 and x > 0.4 a signal at

-157 ppm has been observed, resulting from potassium aluminium hexafluoride

(K3AlF6*0.5H2O) (Harris and Jackson, 1991). This is in agreement with the results

reported by Fechtelkord et al. (2003a) who also observed the formation of potassium

aluminium hexafluoride in F-rich phlogopites synthesised at 800 °C. However, this

phase was already present in their samples with x > 0.1.

It should be noted that the relative intensity of the AlF63-signal in the spectra

shown does not correspond to the true amount of this impurity in the sample. Due to

* * *

F-Mg3

F-Mg Al2

AlF63-

a)

* ** *ppm-220-200-180-160-140

*

F-Mg3

F-Mg Al2

AlF63-

b)

ppm-220-200-180-160-140

Figure 4.25. Comparison of 19F MAS NMR spectra of phlogopites of composition xnom = 0.8, y = 0.5 synthesised

at 600 (a) and at 800 °C (b).

4.3. Ordering of cations and anions in the octahedral sheets of phlogopite

101

the much faster spin-lattice relaxation of this compound, its intensity is highly

overestimated. Nevertheless, a comparison of samples synthesised at both

temperatures is possible qualitatively. In Figure 4.25 two such spectra for an

exemplary composition of xnom = 0.8, y = 0.5 are shown. A strong decrease of AlF63-

formation is clearly visible for the lower synthesis temperature.

4.3.4. J-formalism and Monte-Carlo simulations

The computations have been performed for F-free compositions only and, thus,

only account for the ordering of cations in the octahedral sheets without considering

the ordering of the OH/F anions. For pure Mg-phlogopite, no ordering is possible as

only Mg-atoms occupy the octahedral sites. As Al is incorporated into the structure, it

is thought to order on the slightly smaller M2 site in natural phlogopite samples

(Cruciani and Zanazzi, 1994; Brigatti et al., 2000). For the synthetic Al-rich phlogopite

samples synthesised in the frame of this work, short-range ordering in a way such

that never two Al-atoms occupy directly neighboured octahedral sites has been found

(see previous chapters).

J = 0.58 eV12

J = 0.14 eV13J = 0.15 eV14

J = 0.06 eV15

Figure 4.26. Comparison of the octahedral intrasheet J-parameters. The error range of the values is ± 0.05 eV.

4. Results and discussion

102

a)

b)

Figure 4.27. Configuration of lowest energy derived by Monte-Carlo simulations for a single octahedral sheet of

phlogopites with composition a) x = 0.25, y = 2.0 (K (Mg2.75Al0.25) (Al1.25Si2.75O10) (OH)2) and b) x = 0.75, y = 2.0

(K (Mg2.25Al0.75) (Al1.75Si2.25O10) (OH)2). Mg-ions are shown in green, Al-atoms in red. Grey bars indicate Mg-Al

neighbour pairs. Only a part of the supercell is shown.

4.3. Ordering of cations and anions in the octahedral sheets of phlogopite

103

Figure 4.28. Configuration of lowest energy derived by Monte-Carlo simulations for a single octahedral sheet of

phlogopite with composition x = 1.0, y = 2.0 (K (Mg2Al) (Al2Si2O10) (OH)2). Mg-ions are shown in green, Al-atoms

in red. Grey bars indicate Mg-Mg neighbour pairs. Only a part of the supercell is shown.

The values obtained for the J -parameters describing ordering in the octahedral

sheets are shown in Figure 4.26. The distribution of J -values looks similar to that of

the tetrahedral sheets: The first parameter, 12J , is much larger than the other

parameters, and the values decrease with increasing distance of the corresponding

sites. However, the second and third value, 13J and 14J , are now equal. 15J , the last

parameter is already approximately zero.

With a value of 0.57 ± 0.05 eV, the 12J -parameter is still smaller than the

corresponding parameter for ordering in the tetrahedral sheets, 1J = 1.07 ± 0.05 eV.

Nevertheless, its influence on the ordering of ions is the same: It is highly

unfavourable for Al-atoms to occupy neighbouring octahedral sites. Instead every Al-

atom is surrounded by six Mg-atoms.

Two examples for the obtained configurations of lowest energy are shown in

Figure 4.27. It becomes clear that the Al-atoms are not distributed equally over the

whole octahedral sheet but clustered in some areas. A division into Al-free clusters

and strongly ordered Al-rich clusters with Mg/[4]Al = 2:1 takes places. Only a few

single Al-atoms can be found in the Mg-rich areas of the structure. When going from

4. Results and discussion

104

low (x = 0.25, a) to high (x = 0.75, b) Al-contents only the relative amounts of Al-free

and Al-rich clusters change. At a maximum composition of x = 1.0 only the ordered

clusters with Mg/[4]Al = 2:1 remain (Figure 4.28). Different configurations of Mg and Al

can be distinguished in the ordered clusters, forming separate domains with point

defects at the domain borders (lower part of Figure 4.28).

To compare the simulation results with the experimental observations, it is useful

to describe the distribution of Mg and Al in the octahedral sheets related to the

environments of OH-groups despite the fact that the latter have not been an ordering

species considered in the simulations. The absolute numbers of hydroxyl-groups

having zero, one, two, and three Al-atoms in their environments is counted, and the

relative amounts of these environments are calculated. These values can be directly

compared to the relative intensities of the corresponding signals obtained from 1H

MAS NMR spectra of F-free phlogopites. The results are given in Table. 4.3.

Both simulations and NMR results are in good agreement, showing an avoidance

of Al-Al neighbouring pairs in the octahedral sheet. It should be noted that in the

simulations, four runs have been performed for each Al-content x, and the numbers

of Al-atoms in the hydroxyl environment have been found to be exactly the same for

each run. This indicates that these numbers are not dependent on composition.

The question of whether the clustering found for the configurations of lowest

energy is actually present in the synthesised phlogopite samples cannot be

answered experimentally using 1H MAS NMR spectroscopy. However, it is possible

to gather additional information on the distribution of protons in the octahedral sheet

Table 4.3. Comparison of relative numbers of H-OMg3 and H-OMg2Al environments determined from MC

simulations and from 1H MAS NMR spectroscopy.

Number of Al-atoms in OH

environments (simulations) [%] Relative 1H MAS NMR signal

intensities [%]

x 0 Al 1 Al 2 Al 3 Al Mg3 Mg2Al MgAl2 Al3

0.25 75 25 0 0 63 37 0 0

0.68 32 68 0 0 30 70 0 0

0.82 18 82 0 0 23 77 0 0

4.3. Ordering of cations and anions in the octahedral sheets of phlogopite

105

when investigating the relationship between the ordering in the two types of sheets

by {1H} → 29Si CPMAS NMR spectroscopy as is described in Chapter 4.4. The

results presented there confirm the observed clustering of Al in the octahedral

sheets.

4. Results and discussion

106

4.4. Relationship between the ordering of ions in the tetrahedral and

in the octahedral sheets of phlogopite

In the previous chapters, it has been shown that both ordering of ions in

tetrahedral and octahedral sheets are dominated by avoidance of Al-O-Al linkages.

Complete long-range order has been found for the hypothetical Al-rich end-member

K (Mg2Al) (Al2Si2O10) (OH,F)2. In tetrahedral sheets, Al and Si atoms alternate in

neighbouring tetrahedra, while in octahedral sheets, Al is always surrounded by six

Mg-atoms as next-nearest-neighbours.

For Al-contents x < 1.0, clustering is present. In the tetrahedral sheets clusters of

Si/[4]Al = 1:1 showing perfect ordering of cations can be separated from disordered

clusters of compositions close to Si/[4]Al = 3:1. In single octahedral sheets a similar

clustering takes place: Nearly all Al is enriched in clusters of composition Mg/[6]Al =

2:1. Hardly any Al is present in the remaining parts of the structure.

The question now arising is whether there is any relationship between the

clustering in both sheets and how both types of clusters are oriented with respect to

each other. To simplify the ordering system, we first focus on F-free compositions.

This reduces the number of ordering schemes to two: Mg/[6]Al and Si/[4]Al. Later on

OH/F ordering will be considered either.

4.4.1. Hydroxyl-phlogopites (y = 2.0)

4.4.1.1. 2D {1H} → 29Si HETCOR CPMAS NMR spectroscopy

2D {1H} → 29Si hetero-nuclear correlation (HETCOR) CPMAS NMR spectroscopy

is an ideal tool to investigate the relationship between ordering in both sheets

because it combines information on the local 1H environment in the octahedral sheet

with that on tetrahedral 29Si environments nearby (see Chapters 2.2.2.2 and 2.2.2.3).

One such 2D {1H} -> 29Si HETCOR NMR spectrum is shown in Figure 4.29 for a

composition of xnom = 0.5. The 29Si signals of the F2 dimension can be assigned to

Si-Al3 (-80 ppm), Si-SiAl2 (-83 ppm), Si-Si2Al (-87 ppm), and Si-Si3 (-91 ppm), as has

been discussed in Chapter 4.2.1. These are correlated to the 1H MAS NMR signals in

the F1-dimension with the H-OMg3 signal being located at about 0.7 ppm and the

H-OMg2Al signal at 2 ppm (see 1D 1H MAS NMR spectra shown in Chapter 4.3.1.1).

4.4. Relationship between the ordering of ions in the tetrahedral and in the octahedral sheets of phlogopite

107

-65 -70 -75 -80 -85 -90 -95

-4

-2

10

8

6

4

2

0

F2 (ppm)

F1

(ppm

)

29Si

1H

Si-Si2Al

Si-SiAl2

Si-Al3

H-OMg3

H-OMg Al2

Figure 4.29. 2D {1H} → 29Si HETCOR CPMAS NMR spectrum of phlogopite with nominal composition

K(Mg2.2Al1.8)(Al1.8Si2.2O10)(OH)2 (xnom = 0.5).

a

bc

Mg/Al

OH/F

O

K+

Si/Al

Figure 4.30. View on the tetrahedral sheets of phlogopite. Every OH-position is co-ordinated by three octahedral

cations which may be either Mg or Al (white arrows). The information on a single OH environment is passed on to

six neighbouring tetrahedra if these are occupied by 29Si (black arrows). Each tetrahedral site has three next-

nearest-neighbours which may be either Si or Al. In this way, the number of Al co-ordinating OH may be

correlated to the amount of Al in the 29Si environment in the tetrahedral sheet.

4. Results and discussion

108

Every hydroxyl-group is surrounded by three cations in the octahedral sheet

(white arrows in Figure 4.30), i.e. either 3 Mg or 2 Mg and 1 Al. On transfer of

magnetisation from proton to 29Si, information on each specific OH environment is

passed on to up to six sites in each of the neighbouring tetrahedral sheets, if they are

occupied by 29Si (black arrows). The tetrahedral cations in return have three next-

nearest-neighbours in the tetrahedral sheet, which may be either Al or other Si-

atoms. Each tetrahedron receives magnetisation from three different OH-sites. In this

way, it is possible to gather information on the amounts of ‘pairs of environments’ in

the structure, e.g., how many H-OMg3 environments are located close to Si-Si3

environments. The corresponding signals in the resulting 2D HETCOR spectrum are

proportional to the relative amounts of ‘pairs of environments’ in the sample.

The intensity of the Si-SiAl2-signal is the same for both the H-OMg3 and the

H-OMg2Al signals, which means that this Si environment can be found next to both

types of H environments in equal amounts. The signal intensity corresponding to the

Si-Al3 ↔ H-OMg2Al environment pair, however, is much higher than the intensity of

the Si-Al3 ↔ H-Mg3 signal. The opposite is true for the Si-Si3 environment where the

Si-Si3 ↔ H-Mg3 signal shows higher signal intensity. Al-rich Si environments in the

tetrahedral sheet are more likely to be found in direct neighbourhood of Al-rich proton

environments in the octahedral sheets. In contrast, 29Si having a lower number of Al-

atoms as next-nearest-neighbours more often are located next to Al-free OH

environments.

Taking into account the clustering of Al in both sheets, the results indicate that Al-

rich clusters in octahedral sheets are directly neighboured to Al-rich clusters in

tetrahedral sheets, and there is a relationship between the ordering patterns in both

types of sheets.

4.4.1.2. J-formalism and Monte-Carlo simulations

MC simulations have been performed to obtain additional information on the

distribution of ions in the phlogopite structure. Interactions between tetrahedral sites

of the same layer package, between tetrahedral sites of adjacent layer packages,

and between neighbouring octahedral and tetrahedral sites have been considered

(Chapter 2.3), and the values obtained for the J -parameters are presented in Figure

4.31.

4.4. Relationship between the ordering of ions in the tetrahedral and in the octahedral sheets of phlogopite

109

J5 J6J11J8 J9 J10

J7 J16 J17 J18 J19

tetrahedralintralayerinteraction

tetrahedralinterlayerinteraction

octahedral-tetrahedralinteraction

Figure 4.31. Values of the tetrahedral intralayer, tetrahedral interlayer, and the octahedral – tetrahedral interaction

parameters obtained from GULP.

The tetrahedral intralayer interaction parameters are positive, but only of low

values indicating that there is a slight Al-Al avoidance between the two tetrahedral

sheets of one layer package. The J -parameters describing the interaction between

tetrahedral sheets of adjacent layer packages are approximately zero. Therefore

interactions affecting more than one layer package do not play any role in cation

ordering.

However, interaction parameter 16J , corresponding to pairs of directly

neighboured tetrahedral and octahedral sites, is highly negative. Therefore, Al-O-Al

linkages are in fact energetically favourable if two different types of polyhedra are

involved. This is in contrast to next-nearest-neighbour-interactions within tetrahedral

and octahedral sheets, where Al-Al-pairs are avoided. The other parameters

describing interaction between octahedral and tetrahedral sheets are negative, too,

meaning that the positive effect of [4]Al and [6]Al being positioned close to each other

is not restricted to directly neighboured sites, but affects a larger area of up to three

or four sites.

For x = 1.0, the resulting configuration of lowest energy shows the ordering

patters described in Chapters 4.2.4 and 4.3.4 for single tetrahedral and octahedral

sheets, respectively. Si/[4]Al are perfectly ordered on alternating sites. Si-O-Si, and

4. Results and discussion

110

a)

b)

Figure 4.32. Details of the configuration of lowest energy obtained from MC simulations for x = 1.0. Only one 2:1

layer package is given. Si, [4]Al, Mg, and [6]Al-atoms are shown in yellow, red, blue, and green, respectively. In the

lower picture, Mg has been omitted for clarity. Grey bars connect pairs of directly neighboured Si- and Al-atoms.

Domains can be distinguished by the different orientation of [4]Al in the lower tetrahedral sheet to [4]Al in the upper

tetrahedral sheet. Two such configurations are marked by white ellipsoids. Some of the domain boundaries are

highlighted by white lines. They are characterised by Si-O-Si and Al-O-Al linkages.

4.4. Relationship between the ordering of ions in the tetrahedral and in the octahedral sheets of phlogopite

111

Al-O-Al linkages are only found as defects at domain boundaries. The domains are

characterised by different configurations of Al-atoms in two tetrahedral sheets of the

same layer package: They are either located directly on top of each other, or Al in the

upper tetrahedral sheet is displaced by one tetrahedron relative to Al in the lower

tetrahedral sheet (Figure 4.32).

Mg and [6]Al are perfectly ordered, too, with Al always being co-ordinated by six

Mg in the neighbouring octahedra. Distinction between separate domains is also

possible by the orientation of octahedral Al with respect to the rings of tetrahedra. [6]Al may occupy two different positions within the hexagonal rings or one position at

the edges of the rings, and all three positions seem to be equally favourable.

At lower Al-contents, the clustering observed for single tetrahedral and octahedral

sheets becomes visible again. An example for the resulting ordering patterns is

shown in Figure 4.33. It is obvious that a coupling between ordering in both of the

tetrahedral sheets belonging to one layer package is present. Al-rich clusters of both

sheets are located close to each other. At the same time octahedral Al can only be

found close to Al-rich tetrahedral clusters due to the highly positive value of J16. In

areas with low amounts of tetrahedral Al hardly any [6]Al is observed.

As a result, all Al incorporated into the phlogopite structure is enriched in clusters

of the hypothetical end-member ‘eastonite’, K (Mg2Al) (Al2Si2O10) (OH)2. The clusters

include both tetrahedral and octahedral sheets of a single layer package. The

remaining part of the structure has a composition close to that of the pure Mg end-

member, K Mg3 (AlSi3O10) (OH)2.

These results are quite opposite to what has been observed in MC simulations of

ordering in the dioctahedral mica phengite (K (Al1.5Mg0.5) (Al0.5Si3.5O10) (OH)2) by

Palin et al. (2003). In phengite, Mg is associated with tetrahedral Al, while Si and

octahedral Al are located close to each other. However, the chemical composition of

phengite differs completely from that of Al-rich phlogopite. The former has a much

higher amount of [6]Al and a lower amount of [4]Al in addition to being dioctahedral.

A possibility to check the agreement between experiment and theory is to try to

reproduce the 2D HETCOR NMR spectra, at least in numbers. The following

procedure has been used to get the correlations: The OH-sites could only be marked

indirectly, as they have been integrated into the structural body, and only the cation

sites have been treated as real sites on which ordering occurs. Nevertheless, it was

4. Results and discussion

112

b)

a)

Figure 4.33. Details of the configuration of lowest energy obtained from MC simulations for x = 0.5. Only one 2:1

layer package is given. Si, [4]Al, Mg, and [6]Al-atoms are shown in blue, red, green, and yellow, respectively. In the

lower picture, Mg has been omitted for clarity. Grey bars indicate Al-O-Si linkages in the tetrahedral sheet.

4.4. Relationship between the ordering of ions in the tetrahedral and in the octahedral sheets of phlogopite

113

possible to count the numbers of Al neighbours for each fictive OH-site (see Figure

4.30). Then for each Si-atom the three closest OH-sites have been investigated and

one number has been added for the corresponding correlations. The absolute

numbers are then over-counted, as every OH-site has more than only one

neighbouring Si-site. However, this over-counting also occurs in the real 2D

HETCOR experiments, as protons transfer magnetisation to all Si-sites within a

certain distance range.

In Figure 4.34, a comparison of the values obtained for a composition of x = 0.68

and the 2D {1H} → 29Si HETCOR MAS NMR spectrum of a phlogopite of the same

composition (i.e., xest = 0.68) is given. There are large discrepancies between

simulation and experiment: The Si-SiAl2 signal intensity is the same for both the

H-Mg2Al and the H-OMg3 environments in the spectrum. The simulation results,

however, give a much higher number of 966 Si-SiAl2/H-OMg2Al environment pairs

compared to only 36 Si-SiAl2/H-OMg3 environment pairs. Also, the number of the

Si-Al3/H-OMg3- and the Si-Si2Al/H-OAlMg2 environment pairs is by far

underestimated. However, the trend in both is the same: Mg-rich OH-sites in the

octahedral sheet prefer Si-rich Si environments as nearest neighbours in the

tetrahedral sheet, while Al-rich environments in both types of sheet tend to be located

next to each other.

-65 -70 -75 -80 -85 -90 -95

-4

-2

10

8

6

4

2

0

F2 (ppm)

F1

(ppm

)

29Si

1H

Si-Si2Al

Si-SiAl2

Si-Al3

H-OMg3

H-OMg Al2

Si-Al3 Si-SiAl2 Si-Si Al2 Si-Si3

H-OMg3

H-OMg Al2

H-OMgAl2

H-OAl3

82 1005 966 825

3920 231 36 63

0 0 0

0 0 0 0

0

Figure 4.34. Comparison of site connectivies obtained from MC simulations for a composition of x = 0.68 (left) to

the 2D {1H} → 29Si CPMAS HETCOR NMR spectrum of a phlogopite with the same estimated Al-content.

4. Results and discussion

114

4.4.2. F-containing phlogopites

In Chapter 4.3 it was shown that F-anions in the octahedral sheets prefer Mg-rich

environments while OH-groups tend to be co-ordinated by two Mg and one Al. At the

same time Al is not distributed equally within the sheets but enriched in clusters

encompassing both tetrahedral and octahedral sheets. This indicates, that hydroxyl-

groups and fluorine anions are also clustered, with OH being located in Al-rich areas

of the structure.

A sophisticated method to test this assumption is {1H} → 29Si contact-time

dependent cross-polarisation (CP) MAS NMR spectroscopy (see Chapter 2.2.2.2). In

contrast to 1D CPMAS NMR spectroscopy, the contact time between both systems is

increased step by step in a number of experiments. The amount of magnetisation

being transferred is a function of the contact time, and the shape of the resulting

curve depends on whether a large proton spin reservoir or an isolated 1H-29Si spin

system is present.

The experiments have been performed on samples synthesised at 600 °C in the

frame of this thesis, but also on samples prepared at 800 °C and already described

by Fechtelkord et al. (2003a,b). For each experiment, two magnetisation curves have

been recorded, one regarding the maximum intensity (i.e., signal height) of the

Si-Si2Al or the Si-SiAl2 signal (depending on sample composition), and the other

considering the overall signal area of all Si-nAl signals. Both types of resulting curves

are shown in Figure 4.35. Additionally, all data are given in Table 6.6 in the

Appendix. A more detailed analysis has been performed on selected samples,

investigating the integrals of all signals separately. It has been found that the shape

of the magnetisation curve does not depend on the specific signal used for analysis.

The experimental curves clearly do not show any sign of oscillatory parts.

Magnetisation is transferred from a large spin reservoir, and thus, OH must also be

clustered in the structure. Considering the strong preference of hydroxyl-groups for a

co-ordination by 2 Mg and 1 Al, it is likely for OH to be found in the Al-rich clusters of

composition x = 1.0.

The data have been fitted according to equation (2.33), and an exemplary fit is

shown in Figure 4.36. The values for the cross-polarisation time THSi and the spin-

lattice relaxation time in the rotating frame, T1ρ, obtained from the fits are given in

4.4. Relationship between the ordering of ions in the tetrahedral and in the octahedral sheets of phlogopite

115

Figure 4.35. Experimental magnetisation curves for signal area (top, all Si-nAl signals) and highest signal intensity

(bottom, Si-Si2Al or Si-SiAl2 signal).

4. Results and discussion

116

Table 4.4. Initial magnetisation M0, cross-polarisation time THSi, spin-lattice relaxation time in the rotating frame

T1ρ, and mean H-Si distance dH-Si obtained from fits of {1H} → 29Si magnetisation curves of Al-rich phlogopites

K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y (F)2-y. The error range of THSi is ± 1.0 s, that of dH-Si ± 0.1 Å.

signal area signal intensity

x y T

[°C] M0

[a.u.] T1ρ [s]

THSi [ms]

dH-Si [Å]

M0 [a.u.]

T1ρ [s]

THSi [ms]

dH-Si [Å]

0.4 1.0 800 4.1·1010 > 2 4.5 3.7 3.6·108 2.4 5.0 3.8

0.5 1.0 800 4.6·1010 > 5 4.5 3.7 4.2·108 > 4 5.0 3.8

0.6 1.0 800 3.6·1010 > 10 4.0 3.6 3.3·108 > 10 4.0 3.6

0.7 1.0 800 4.0·1010 > 10 4.0 3.6 4.4·108 > 10 4.0 3.6

0.8 1.0 800 2.5·1010 > 5 4.5 3.7 2.8·108 > 4 5.0 3.8

0.4 1.5 800 4.0·1010 > 7 3.5 3.6 3.0·108 > 5 2.5 3.4

0.6 1.5 800 4.0·1010 > 10 3.0 3.5 3.5·108 > 10 2.5 3.4

0.8 1.8 800 3.6·1010 > 10 2.5 3.4 3.8·108 > 10 2.5 3.4

0.5 1.6 600 4.1·1010 > 15 2.5 3.4 3.1·108 > 15 2.5 3.4

0.7 1.8 600 4.1·1010 > 5 2.5 3.4 2.8·108 > 5 2.5 3.4

Table 4.4. The mean H-Si distance has been calculated from THSi, using equation

(2.19). The results obtained from signal area fits and signal intensity fits are roughly

in the same range. With increasing OH-content of the nominal gel composition, the

cross-polarisation time decreases from about 5 ms to 2.5 ms. Therefore, also the

estimated distance between protons and Si-atoms decreases slightly from 3.7 Å to

3.4 Å. Tateyama et al. (1974) reported slightly shorter H-Si distances of 3.18 Å,

3.20 Å, and 3.22 Å for a hydroxyl-phlogopite.

Figure 4.36. Magnetisation curve derived from the highest intensities of the Si-Si2Al signal for phlogopite of

nominal composion xnom = 0.7, y = 1.0 (synthesis temperature T = 800 °C). The solid line represents a fit to the

data according to equation (2.33).

4.4. Relationship between the ordering of ions in the tetrahedral and in the octahedral sheets of phlogopite

117

Figure 4.37. Experimental magnetisation curves of F-rich phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y derived

from the signal area of all Si-nAl signals. Top: Whole contact time range. Bottom: Detail of low contact times.

4. Results and discussion

118

Table 4.5. Initial magnetisation M0, cross-polarisation time TFSi, spin-lattice relaxation time in the rotating frame

T1ρ, and mean F-Si distance dF-Si obtained from fits of {19F} → 29Si magnetisation curves for F-rich phlogopites

K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y (F)2-y. The error range of THSi is ± 1.0 s, that of dH-Si ± 0.1 Å.

signal area signal intensity

x y M0

[a.u.] T1ρ [ms]

TFSi [ms]

dF-Si [Å]

M0 [a.u.]

T1ρ [ms]

TFSi [ms]

dF-Si [Å]

0.1 0.2 7.9·1012 41 40 5.2 9.9·1010 41 40 5.2

0.0 0.5 4.9·1012 37 37 5.2 2.9·1010 39 37 5.2

0.3 0.5 2.8·1010 980 6.7 3.9 3.3·108 >1000 6.1 3.8

0.6 0.5 4.8·1010 300 8.6 4.0 4.2·108 310 8.6 4.0

0.8 0.8 4.5·1010 >5000 8.6 4.0 4.2·108 - 8.6 4.0

0.7 1.0 3.2·1010 >5000 10.0 4.2 2.8·108 >1000 10.7 4.2

For all samples, the spin-lattice relaxation time T1ρ is in the range of seconds, i.e.

infinite in the frame of the CPMAS experiments. This means that nearly no motion

processes can be found in the lattice at a frequency of 33 kHz (the nutation

frequency).

In the same way {19F} → 29Si contact-time dependent CPMAS NMR

spectroscopic experiments have been performed on samples synthesised at 600 °C.

The resulting magnetisation curves derived from the overall Si-nAl signal area are

shown in Figure 4.37, and the data are given in Table 6.7 in the Appendix. Most of

the curves are very similar to those obtained from {1H} → 29Si CPMAS NMR

experiments, although the quality of the data is not as good. Especially at very long

contact times the data points strongly scatter due to fluctuations in the tube

transmitter power. Nevertheless, values for T1ρ and TFSi could be obtained for some

of the samples. The curves also show very long to infinite spin-lattice relaxation times

T1ρ (Table 4.5). The fitted cross-polarisation times TFSi are much longer than has

been observed for 1H, leading to slightly longer F-Si distances of about 4 Å compared

to dH-Si ~ 3.5 Å.

The magnetisation curves of the most F-rich and Al-poor samples (x = 0.1,

y = 0.2; x = 0.0, y = 0.5), however, exhibit completely different characteristics. The

increase of magnetisation at low contact times is much slower than for the other

samples (Figure 4.37 bottom), due to long cross-polarisation times of about 40 ms.

This corresponds to long F-Si distances of about 5 Å. In structures of fluoro-

4.4. Relationship between the ordering of ions in the tetrahedral and in the octahedral sheets of phlogopite

119

phlogopites reported by Takeda and Morosin (1975) and McCauley et al. (1973) this

distance is much shorter with 3.53 – 3.54 Å.

At the same time, a decrease of magnetisation has been observed for high

contact times, with T1ρ being only slightly longer or equal to TFSi. Motion processes

with frequencies of about 45 kHz (the nutation frequency) must be present in the

lattice responsible for a loss of energy.

The discrepancy between both {19F} → 29Si CPMAS data sets is very large (~1 Å)

and cannot be explained solely by structural changes due to varying compositions of

the phlogopites. Nevertheless, the results indicate that F-rich, Al-poor samples show

completely different structural characteristics than more OH- and Al-rich phlogopites.

The lack of oscillatory parts in the {19F} → 29Si CPMAS magnetisation curves

indicated that F is also clustered in the structure. 1D {19F} → 29Si CPMAS NMR

experiments should be able to relate this result to the clustering of cations in the

tetrahedral sheets. In Chapter 4.3 it has been shown that the amount of F being co-

ordinated by Mg2Al is always very low compared to the whole amount of F in the

structure. This means nearly all of the magnetisation transferred from 19F to 29Si

nuclei comes from F-Mg3 environments. Therefore, 1D {19F} → 29Si CPMAS NMR

spectra should reflect the local composition and distribution of Si environments next

to F-rich clusters.

The fit parameters obtained from 1D {19F} → 29Si CPMAS NMR spectra of

selected phlogopite samples are given in Table 4.6. For comparison, the

corresponding parameters retrieved from 29Si MAS NMR spectra are also shown. For

the three last samples, the 19F environment has been found to be more Si-rich than

the average composition because the Al-content estimated from CPMAS data is

slightly lower than that calculated from 29Si MAS NMR spectra. In many cases the

lower Al-content is also visible from a shift of the Si-nAl signal positions to more

shielded values.

In principle, the sample with composition xnom = 0.4,y = 0.5 shows the same

trend. For a contact time of 3 and 7 ms, xest determined from CPMAS NMR is lower

than the overall xest. However, for a contact time of 5 ms the estimated Al-content is

equal to the average value.

4. Results and discussion

120

Tab

le 4

.6.

Fit

para

met

ers

obta

ine

d fr

om 1

D {

19F

} →

29S

i C

PM

AS

NM

R s

pect

ra o

f ph

logo

pite

s K

(M

g 3-xA

l x) (

Al 1

+xS

i 3-x

O1

0)

(OH

) y F

2-y r

ecor

ded

with

diff

eren

t co

ntac

t tim

es.

For

co

mpa

rison

fit

para

met

ers

for

29S

i MA

S N

MR

spe

ctra

are

als

o g

iven

. P

os.

= p

ositi

on,

FW

HM

= f

ull w

idth

at

half

max

imum

, F

= r

elat

ive

sign

al a

rea.

The

app

roxi

mat

e er

ror

rang

e

for

the

sign

al a

rea

is ±

2 %

.

Si-O

-Si 3

A

[%]

12

9 8 7 7 12

9 11

6 8 9 9 9 6 12

8 8 7 12

9 11

FW

HM

[ppm

]

2.9

2.4

2.4

2.4

2.8

2.6

2.6

2.6

2.8

2.4

2.4

2.4

2.4

2.8

2.4

2.4

2.4

2.8

2.8

2.8

2.8

pos.

[ppm

]

-92.

4 -9

1.5

-91.

2

-90.

8

-90.

5

-92.

4

-91.

5

-91.

3

-91.

3

-91.

5

-91.

6

-91.

5

-91.

5

-91.

4

-91.

0

-91.

1

-91.

0

-90.

6

-91.

4

-91.

4

-91.

5

Si-O

-Si 2

Al

A

[%]

47

43

41

41

47

48

46

49

44

47

47

46

46

40

43

45

45

27

38

38

36

FW

HM

[ppm

]

2.7

2.3

2.3

2.3

2.5

2.3

2.3

2.3

2.6

2.3

2.3

2.3

2.3

2.5

2.3

2.3

2.3

2.5

2.5

2.5

2.5

pos.

[ppm

]

-88.

4 -8

7.8

-87.

6

-87.

5

-86.

9

-88.

9

-87.

7

-87.

8

-87.

8

-88.

0

-88.

0

-88.

0

-88.

0

-87.

4

-87.

5

-87.

4

-87.

5

-87.

2

-87.

6

-87.

7

-87.

6

Si-O

-SiA

l 2

A

[%]

32

43

43

43

42

37

37

37

44

40

40

40

40

45

38

41

41

43

40

41

42

FW

HM

[ppm

]

2.6

2.4

2.4

2.4

2.6

2.5

2.4

2.4

2.6

2.4

2.4

2.4

2.4

2.6

2.4

2.4

2.4

2.5

2.5

2.5

2.5

pos.

[ppm

]

-84.

6 -8

4.3

-84.

1

-84.

1

-83.

3

-85.

4

-84.

2

-84.

3

-84.

3

-84.

5

-84.

6

-84.

5

-84.

5

-83.

8

-84.

0

-84.

0

-84.

0

-83.

6

-84.

1

-84.

1

-84.

1

Si-O

-Al 3

A

[%]

7 5 9 9 4 3 8 4 6 5 4 5 5 8 7 6 6 23

10

11

11

FW

HM

[ppm

]

2.9

2.4

2.4

2.4

2.8

2.6

2.6

2.6

2.8

2.4

2.4

2.4

2.4

2.8

2.4

2.4

2.4

2.6

2.6

2.6

2.6

pos.

[ppm

]

-80.

4 -8

0.7

-80.

6

-81.

0

-80.

0

-81.

7

-81.

1

-81.

2

-80.

8

-81.

2

-81.

4

-81.

0

-81.

2

-80.

2

-80.

4

-80.

6

-80.

5

-79.

9

-80.

8

-80.

5

-80.

6

x est

0.24

0.30

0.35

0.36

0.29

0.22

0.30

0.24

0.33

0.28

0.27

0.28

0.28

0.37

0.27

0.30

0.30

0.51

0.32

0.36

0.35

cont

act

time

[ms]

- 3 5 7 - 3 5 7 - 3 5 7 9 - 3 5 7 - 3 5 7

y

0.0

0.5 0.5 0.8 0.1

x

0.8

0.4

0.7

0.4

0.7

4.4. Relationship between the ordering of ions in the tetrahedral and in the octahedral sheets of phlogopite

121

Only for the F-free composition (y = 0.0) an opposite trend has been observed

with the 19F environment showing a higher Al-content. So far, no explanation has

been found for this contradictory result.

Nearly all samples also showed a decrease in FWHM of the CPMAS signals

compared to 29Si MAS NMR signals. The latter contain information on all the Si

environments throughout the structure. Si-O-Al bond length and bond angles slightly

differ for Si-atoms in both types of clusters leading to broad 29Si MAS NMR signals.

In contrast, in {19F} → 29Si CPMAS NMR experiments only Si-atoms in one type of

cluster are considered. These environments are more homogeneous resulting in a

smaller signal width.

4. Results and discussion

122

4.5. 27Al MAS and 27Al 3Q-MAS NMR spectroscopy

In addition to the experiments described before 27Al MAS and MQMAS spectra

have been recorded. These are usually difficult to interpret because 27Al belongs to

the group of quadrupolar nuclei, and signals are expected to be broadened due to

quadrupolar interaction in addition to other line-broadening effects. Nevertheless,

additional information may be obtained on the Al environments in the octahedral and

tetrahedral sheets of phlogopite. Moreover, apart from the powder X-ray diffraction

patterns discussed in Chapter 4.7 27Al MAS and MQMAS spectra are the only way to

obtain information on Al2O3, the most prominent impurity phase in the samples under

investigation.

Figure 4.38 shows a comparison of 27Al MAS NMR spectra of phlogopite samples

with different Al- and OH-contents. The signals at about 60 to 70 ppm result from

tetrahedrally co-ordinated Al, while octahedrally co-ordinated Al gives rise to signals

in the range of 0 – 20 ppm (Müller et al., 1981; Lipsicas et al., 1984). The signals are

broadened due to distributions of chemical shifts and electric field gradients, dipolar

coupling, and low crystallinity of the samples, and hardly any features of typical

quadrupolar patterns can be distinguished.

All spectra are dominated by a strongly asymmetric signal at δ(27Al) = 71 ppm.

This signal is due to [4]Al which substitutes for Si in the tetrahedral sheets of

phlogopite. It has been fitted using a simple version of the Czjzek distribution model

which is implemented in the DMFit software and takes into account a distribution of

chemical shifts (Massiot et al, 2002). The position of δ(27Al) = 69.5 to 72.6 ppm is in

agreement with the findings of Circone et al. (1991) and Fechtelkord et al. (2003b). In

consistence with Woessner (1989), the [4]Al signal position shifts to less-shielded, i.e.

more-positive values with increasing substitution of Si by Al due to an increase of

distortion of the tetrahedral layer (Figure 4.38). The quadrupolar coupling constant

CQ is in the range of 2.4 to 2.9 MHz, the full width at half maximum of the Gaussian

distribution of chemical shifts is between 3.2 and 4.6 ppm, and no trends with

changing sample composition have been observed. The asymmetry parameter η has

not been taken into account in the distribution model and thus could not be

determined.

4.5. 27Al MAS and 27Al 3Q-MAS NMR spectroscopy

123

y =

1.0

y =

1.2

y =

1.4

y =

1.6

y =

1.8

x =

0.2

x =

0.4

x =

0.6

x =

0.8

-40

804

00

ppm

-40

8040

0p

pm-4

080

40

0pp

m

-40

80

40

0p

pm

-40

804

00

ppm

-40

8040

0p

pm

-40

8040

0p

pm

-40

80

40

0p

pm

-40

8040

0p

pm

-40

8040

0

ppm

-40

80

400

-40

80

40

0

ppm

ppm

-40

8040

0-4

080

400

ppm

ppm

**

**

**

*

* ***

**

**

-40

8040

0p

pm-4

080

400

ppm

-40

804

00

ppm

*

*

Figure 4.38. Comparison of 27Al MAS NMR spectra of phlogopites of different compositions

K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y. Spinning sidebands are marked by asterisks.

4. Results and discussion

124

uncertainty:

[4] A

l sig

nal

po

siti

on

[p

pm

]

Figure 4.39. Plot of the phlogopite [4]Al signal position against the estimated Al-content xest of phlogopites with

different F-contents y.

The phlogopite [4]Al signal is highly asymmetric, a phenomenon first discussed by

Woessner (1989) for Al in the sheets of several clay minerals. The Al environment is

not constant throughout the sheets but changes slightly, giving rise to a distribution of

chemical and quadrupolar shifts.

With increasing Al-content of the samples the [6]Al signal intensity increases and

at least two separate signals can be distinguished. According to Fechtelkord et al.

(2003b) and Circone et al. (1991), the phlogopite [6]Al signal is expected to be

positioned at δ(27Al) = 10 ppm, and another signal at δ(27Al) = 16 ppm should result

from corundum (α-Al2O3). Fechtelkord et al. (2003a,b) reported an additional signal at

δ(27Al) = 5 ppm for F-rich compositions of y = 0.5 which should be due to smaller

amounts of potassium aluminium hexafluoride (K3AlF6*0.5H2O). However, these

signals cannot be resolved in the spectra shown in Figure 4.38. For this reason, 27Al

MQMAS NMR experiments have been performed. In these, the F2-dimension is

correlated to the F1-dimension in which only the isotropic parts of chemical and

second-order quadrupolar shifts are left (see Chapter 2.2.2.4).

4.5. 27Al MAS and 27Al 3Q-MAS NMR spectroscopy

125

x = 0.8y = 0.5

nom

0 -202060 4080 -40

F2 [ppm]

0

10

20

30

40

50

60

F1

[ppm

]

x = 0.6y = 1.4

nom

x = 0.5y = 1.6

nom

0

10

20

30

40

50

60

F1

[pp

m]

0 -202060 4080 -40

F2 [ppm]

0

10

20

30

40

50

60

F1

[ppm

]

0 -202060 4080 -40

F2 [ppm]

0 -202060 4080 -40

F2 [ppm]

x = 0.8y = 1.6

nom

0

10

20

30

40

50

60

F1

[pp

m]

x = 1.6y = 2.0

nom

0 -202060 4080 -40

F2 [ppm]

0

10

20

30

40

50

60

F1

[ppm

]

*

*

*

x = 0.8y = 2.0

nom

0 -202060 4080 -40

F2 [ppm]

0

10

20

30

40

50

60

F1

[ppm

]

*

Figure 4.40. 27Al MQMAS NMR spectra of phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y. Spinning sidebands

are marked by asterisks. The F1-axis has been labelled according to the C3a-convention (Amoureux and

Fernandez, 1998; Millot and Man, 2002)

4. Results and discussion

126

Even in the 27Al MQMAS NMR spectra the two [6]Al signals are difficult to

separate due to the very similar shifts in the F1-dimension (Figure 4.40), and

information on the quadrupolar parameters and thus on structural information cannot

be obtained. One of the two signals shows a large distribution of chemical shifts due

to slight changes in the Al environment throughout the whole structure, but no

evidence can be given to which phase is responsible for this signal.

To gather additional information on the position and the quadrupolar parameters

of the signals, high-field experiments have been performed leading to a significant

Table 4.7. NMR parameters obtained from 27Al MAS NMR spectra of phlogopite samples

K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)2 recorded at 17.6 T (27Al Larmor frequency = 195.28 MHz).

δ(27Al) FWHM CS1) CQ η

xnom [ppm] ± 0.5 3) [ppm] ± 1.03) [MHz] ± 0.33) ± 0.23

0.4 73.0 - 3.7 0.6

17.3 7.6 5.7 -

12.8 - 4.8 0.3

0.8 74.4 - 3.7 0.6

65.8 - 7.62) -

15.4 7.1 6.4 -

12.1 - 4.7 0.3

1.0 78.3 - 3.82) -

74.5 - 3.3 0.6

67.5 - 3.52) -

15.1 6.9 5.4 -

11.8 - 4.7 0.3

1.2 77.7 - 2.12) -

74.4 - 3.1 0.6

67.6 - 4.02) -

15.7 6.9 5.4 -

11.7 - 4.4 0.3

1.6 73.0 - 3.7 0.6

17.3 7.6 5.7 -

12.8 - 4.8 0.3

1. FWHM CS = Full width at half maximum of the Gaussian chemical shift distribution

2. For signals fitted with Lorentzian lines only, the full width at half maximum (FWHM) is given instead.

3. Error ranges have been estimated by changing the quadruplar coupling parameters in the fit function

manually until a distinct change of χ2 took place.

4.5. 27Al MAS and 27Al 3Q-MAS NMR spectroscopy

127

020406080 ppm

x = 0.4nom

020406080 ppm

x = 0.8nom

020406080 ppm

x = 1.0nom

020406080 ppm

x = 1.2nom

020406080 ppm

x = 1.6nom

Figure 4.41. Comparison of 27Al MAS NMR spectra of phlogopite samples with nominal composition

K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y recorded at 17.6 T.

4. Results and discussion

128

narrowing of the signal full width at half maximum (FWHM) due to lower quadrupolar

interaction. 27Al MAS NMR spectra of samples with compositions y = 2.0 and

xnom = 0.4, 0.8, 1.0, 1.2 and 1.6 were recorded by the group of Dr. Jürgen Haase at

the Department of Interface Physics and the Magnet-Resonanz-Zentrum of the

University of Leipzig at a field strength of 17.6 T (proton Larmor frequency =

750 MHz, 27Al Larmor frequency = 195.28 MHz). These spectra are shown in Figure

4.41, and the fit parameters are given in Table 4.7.

The octahedral region of the spectra is made up by a large, broad signal

(CQ = 5-7 MHz) at δ(27Al) = 15-17 ppm, showing a strong distribution of chemical

shifts, and by a smaller signal at δ(27Al) = 12.0 ppm with a lower CQ of about 4-5 MHz

and an asymmetry parameter η of 0.3. However, additional smaller signals may still

be present. A comparison with Circone et al. (1991) and Fechtelkord et al. (2003b)

shows that the signal at δ(27Al) = 12 ppm should result from octahedrally co-ordinated

Al in the sheets of phlogopite, while the larger signal should result from an Al-oxide

component.

The position of the phlogopite [4]Al signal now ranges from δ(27Al) = 73.0 to

74.5 ppm while in the low field spectra δ(27Al) was in the range of 70.5 to 72.5 ppm. A

distribution of electric field gradients is not visible anymore. This results from the

lower quadrupolar interaction at higher field strengths. One or two smaller signals

also appear in the tetrahedral region between 78 and 65 ppm which could not be

observed at a field of 9.34 T.

Four more samples of compositions xnom = 0.8/y = 0.5, xnom = 0.8/y = 1.0,

xnom = 0.4/y = 1.6, and xnom = 1.2/y = 1.8 have been investigated by Dr. Ulrike

Werner-Zwanziger, Dr. Josef Zwanziger and Dr. Michael Fechtelkord at the NMR-3 of

the Chemistry Department at Dalhousie University at a field strength of 16.45 T

(proton Larmor frequency of 700 MHz, 27Al Larmor frequency of 182.47 MHz). 27Al

MAS NMR as well as 27Al MQMAS NMR spectra were recorded for all compositions.

To allow for a better distinction between signals resulting from phlogopite and

Al-oxide impurity phases, a sample of corundum (α-Al2O3) has been analyzed, too.

The industrial sample (Code 1236, Baker & Adamson Products, Gen. Chem. Div.,

Allied, Chem. Corp, Morristown, N.J.) has been heated in the oven at 1050-1100 °C

for 2.5 hours before the 27Al MAS and MQMAS NMR experiments.

4.5. 27Al MAS and 27Al 3Q-MAS NMR spectroscopy

129

Again, the [6]Al signals are not well resolved in the phlogopite spectra

(Figure 4.42). A comparison of the 1D MAS spectra with that of Al2O3 shows that only

the most F-rich sample (xnom = 0.8, y = 0.5) has a strong Al2O3 signal similar to

corundum. For this sample the signal resulting from octahedral Al in phlogopite is

also much smaller than for the other samples. Hence, this F-rich phlogopite must

have incorporated less Al into its crystal structure, and as a result higher amounts of

Al2O3 formed during synthesis.

In the Al2O3 spectrum more signals than the main corundum resonance are

visible: a shoulder at less-negative ppm values from the main signal consisting of at

least one additional signal, and a broad signal of low intensity in the tetrahedral

-30-20-10110 100 90 80 70 60 50 40 30 20 10 0 ppm

x = 0.4y = 1.6

nom

x = 0.8y = 1.0

nom

x = 0.8y = 0.5

nom

x = 1.2y = 1.8

nom

Al O2 3

Figure 4.42. Comparison of 27Al MAS NMR spectra of phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)2 F2 and an Al-

oxide sample recorded at a field strength of 16.45 T (27Al Larmor frequency 182.47 MHz).

4. Results and discussion

130

region of the spectrum. The 27Al MQMAS spectrum shown in Figure 4.43 reveals

three separate signals in the octahedral and up to four small signals in the tetrahedral

region.

When comparing the 1D MAS NMR spectrum to the F2-projection of the MQMAS

spectrum it becomes obvious that the intensity of the main corundum signal at

δiso = 18 ppm has decreased drastically (Figure 4.43). With δ(F2) = 12 ppm and

δ(F1) = 31 ppm it is positioned slightly away from the line indicating an isotropic

environment without quadrupolar interaction, and it has a large quadrupolar

interaction parameter CQ of 5-6 MHz.

The other two octahedral signals at δiso = 15 ppm, δ(F2) = 12 ppm,

δ(F1) = 23 ppm and δiso = 11 ppm, δ(F2) = 9 ppm, δ(F1) = 16 ppm, respectively, are

closer to the diagonal line, and their CQ is only in the range of 3-4 MHz, indicating

that these signals correspond to Al environments of higher symmetry.

ppm

100 80 60 40 20 0 ppm

160

140

120

100

80

60

40

20

0

100 90 80 70 60 50 40 30 20 10 0 ppm

27Al MAS NMR spectrum:

16.4 ppm

22.9 ppm

31.4 ppm

112.8 ppm

123.5 ppm

59.3 ppm

99.9 ppm

F2-projection

F1

-projection

Figure 4.43. 27Al MAS and MQMAS spectra of Al2O3. The 27Al MAS NMR spectrum is shown on top of the F2-

projection of the 27Al MQMAS spectrum. In the left part slices parallel to the F2-axis of the MQMAS spectrum are

shown, and the F1-shifts at which they were taken are given. Labelling of the F1-axis has been done following the

Cz-convention (Millot and Man, 2002) The diagonal line in the MQMAS spectrum indicates positions resulting from

Al environments of high symmetry. In these, no electric field gradient influences the nucleus and thus the signal

shift is only made up by the chemical shift.

4.5. 27Al MAS and 27Al 3Q-MAS NMR spectroscopy

131

Table 4.8. Parameters obtained from 27Al MQMAS NMR spectra of phlogopites and Al2O3, recorded at 16.45 T.

iso )2(F )1(F QC

)( 1obs

zG

xnom y [ppm] [ppm] [ppm] [MHz]

1.2 1.8 9.7 ± 1.4 6.4 ± 0.3 16.4 ± 0.6 4.3 ± 0.4

13.5 9.1 22.9 5.0 ± 0.4

16.5 9.4 29.3 6.3 ± 0.3

71.6 70.9 102.1 2.0 ± 0.9

77.3 76.0 110.6 2.7 ± 0.7

0.4 1.6 9.8 ± 1.7 6.7 ± 0.3 16.4 ± 0.8 4.1 ± 0.7

16.0 8.0 29.3 6.6 ± 0.3

71.0 69.3 102.1 3.1 ± 0.5

0.8 1.0 10.2 ± 1.4 7.8 ± 0.3 16.4 ± 0.6 3.6 ± 0.5

14.0 10.4 22.9 4.5 ± 0.4

16.7 9.9 29.3 6.1 ± 0.3

71.3 70.0 102.1 2.7 ± 0.7

77.5 76.5 110.6 2.4 ± 0.8

0.8 0.5 9.8 ± 1.7 6.7 ± 0.3 16.4 ± 0.8 2.8 ± 0.8

71.2 69.8 102.1 4.1 ± 0.5

Al2O3 10.5 ± 3.6 8.6 ± 0.3 16.4 ± 1.6 3.3 ± 1.5

14.5 11.8 22.9 3.9 ± 1.1

18.3 11.8 31.4 6.0 ± 0.7

39.4 35.3 59.3 4.8 ± 0.9

69.4 67.4 99.9 3.3 ± 1.4

- - 112.3 -

81.1 70.9 123.5 7.5 ± 0.6

At least two smaller signals at δiso = 69.4 ppm and δiso = 81.1 ppm, respectively,

result from tetrahedrally co-ordinated Al. Moreover, another signal with very small

intensity positioned at δiso = 39 ppm indicates that the sample contains small

amounts of [5]Al, too.

It can be concluded that this Al-oxide is characterised by a high degree of

structural disorder. This might also be the case for the Al2O3 impurity phase of the

phlogopite samples, especially if one considers the comparatively low synthesis

temperature of 600 °C that could have prevented a crystallisation of pure corundum.

The quadrupolar parameters determined from the MQMAS spectra of the

phlogopite samples are given in Table 4.8, together with those of Al2O3. Figure 4.44

shows the corresponding spectra.

4. Results and discussion

132

ppm

100 80 60 40 20 0 ppm

150

100

50

0

ppm

100 80 60 40 20 0 ppm

150

100

50

0

100 80 60 40 20 0 ppm

150

100

50

0

100 80 60 40 20 0 ppm

150

100

50

0

x = 1.2y = 1.8

nom x = 0.4y = 1.6

nom

x = 0.8y = 0.5

nomx = 0.8y = 1.0

nom

Figure 4.44. 27Al MQMAS NMR spectra of phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y recorded at 16.45 T

(27Al Larmor frequency 182.42 MHz). The F1-axis has been labelled according to the Cz-convention (Millot and

Man, 2002).

The phlogopite [6]Al signal is located at δiso = 10, δ(F2) = 6-8 ppm and

δ(F1) = 16 ppm. CQ shows a large variation between 2.8 MHz for the F-rich

composition and 4.3 MHz for the OH-rich composition. This is in well agreement with

the results of high-field 27Al MAS spectra of OH-phlogopites investigated at Leipzig

(Table 4.7), which showed a quadrupolar coupling constant CQ of 4.4 to 4.8 MHz.

This is in contrast to the findings of Fechtelkord et al. (2003b) where CQ of F-rich

samples has been much larger (CQ = 5.35 MHz for y = 0.5) than that of OH-rich

samples (CQ = 4.78 MHz for y = 1.8). The dependence on Al-content reported by the

same authors has not been observed in this study.

Similar to corundum, for xnom = 0.8, y = 0.5 the large Al2O3 signal at δiso = 18 ppm

has disappeared in the MQMAS spectrum, hence this sample only shows signals

resulting from Al in phlogopite. In the three other phlogopite spectra, a signal similar

to the one positioned at δiso = 16 ppm in the corundum spectrum has been observed.

4.5. 27Al MAS and 27Al 3Q-MAS NMR spectroscopy

133

δiso, δ(F2), δ(F1), and CQ are roughly in the same range and therefore, the Al

environment in both structures must be similar.

For high Al-samples with xnom = 0.8, y = 1.0 and xnom = 1.2, y = 1.8, respectively,

another signal at δiso = 14 ppm has been observed which is in well agreement with

another one of the Al2O3 signals indicating that this signal is also due to an Al2O3

environment. In agreement with the observations made for the spectra of the OH-

phlogopites described above, in two of these samples another phase containing

tetrahedral Al has been found, and the quadrupolar parameters of this phase could

be determined. However, these parameters are not in agreement with any of the

tetrahedral signals observed in the spectrum of corundum.

4. Results and discussion

134

4.6. 17O MAS and 17O 3Q-MAS NMR spectroscopy

17O MAS and MQMAS experiments have been performed on selected

compositions in order to gather information on the O environment in the structure. It

is of special interest to clarify the presence of Al-O-Al linkages in the tetrahedral

sheets of phlogopite. In 17O MAS and MQMAS spectra, the corresponding signals

should be well distinguishable from signals resulting from Si-O-Al or Si-O-Si

environments.

Due to the very low natural abundance of 17O (0.037 %) this nucleus was

enriched in the sample by using isotopically enriched H2O for hydrothermal synthesis.

Water containing 75-80 at% 17O (Sigma-Aldrich) was necessary to obtain well

resolved 17O MAS and MQMAS spectra.

In Figure 4.45, 17O MAS NMR spectra of phlogopites with xnom = 0.5, y = 0.5 and

xnom = 0.5, y = 1.0, respectively, are shown. Both spectra look very similar, only a

narrow signal at about 70 ppm decreases drastically when going from high (y = 0.5)

to intermediate F-contents (y = 1.0). Several signal components are distinguishable,

but quadrupolar broadening and signal overlapping prevent a detailed analysis of the

spectra.

x = 0.5y = 0.5

nom

x = 0.5y = 1.0

nom

-60-40-20100 80 60 40 20 0

Figure 4.45. 17O MAS NMR spectra of 17O enriched phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y.

4.6. 17O MAS and 17O 3Q-MAS NMR spectroscopy

135

80 70 60 50 40 30 20 10 0

10

20

15

30

35

25

45

40

F2 (ppm)

F1

(ppm

)a)

80 60 40 20 0 ppm80 60 40 20 0

F2 (ppm)

F1

(ppm

)

b)[6] [6]Al-O- Al

Si-O- Al[4]

Si-O-Si

(Si,Al)-O- (Mg,Al)[6]

Figure 4.46. 17O MQMAS spectra of phlogopite with composition xnom = 0.5, y = 1.0. a) Spectrum recorded at

9.34 T (F1-axis labelled according to the C3a-convention; Amoureux and Fernandez, 1998; Millot and Man, 2002).

b) Spectrum recorded at 16.45 T.

4. Results and discussion

136

Therefore, 17O MQMAS experiments at a field strength of 9.34 T have been

performed. In these experiments, the F2-dimension is correlated to the F1-dimension

in which only the isotropic parts of chemical and second-order quadrupolar shifts are

left (Chapter 2.2.2.4).

Moreover, 17O MQMAS high-field spectra have been recorded by Dr. Ulrike

Werner-Zwanziger, Dr. Josef Zwanziger and Dr. Michael Fechtelkord at the NMR-3 of

the Chemistry Department at Dalhousie University at a field strength of 16.45 T

(proton Larmor frequency of 700 MHz, 17O Larmor frequency of 94.94 MHz). The

spectra obtained at both fields are shown in Figure 4.46.

Four signals are already distinguishable at low field, and another fifth signal has

been observed at high field strength. Assignment of these signals to oxygen

environments in the structure is difficult. So far only few 17O NMR studies of clay

minerals have been reported. Signals are supposed to result from the hydroxyl group

(H-O-[6](Mg,Al) environments), from basal oxygen atoms of the tetrahedral sheet

([4](Si,Al)-O-[4](Si,Al) environments), from apical oxygen atoms connecting tetrahedral

and octahedral sheets ([4](Si,Al)-O-[6](Mg,Al) environments), and from the aluminium

oxide impurity phase ([6]Al-O-[6]Al environments). Except for the hydroxyl-group all of

these environments are very similar, and it can be expected that the quadrupolar

parameters do not deviate much. Very well resolved spectra are necessary for a

reasonable distinction between all those signals.

A fit of the spectrum for xnom = 0.5, y = 1.0 is shown in Figure 4.47. Signals have

been fitted manually, considering the MQMAS spectra and data available in the

literature so far. The quadrupolar coupling parameters obtained from the fits are

given in Table 4.9. However, it should be noted that the results are not unique and

other signal parameters may be possible, too.

According to van Eck et al. (1999), the signal corresponding to hydroxyl-groups

should not be observable in 3QMAS spectra due to a short spin-lattice relaxation

time T1. Moreover, this signal is supposed to have a very high CQ of 6 – 7 MHz (van

Eck et al., 1999; Lee et al., 2003a). A slightly lower quadrupolar coupling constant of

5 to 6 MHz has been found in this study.

Lee et al. (2003a,b) performed 17O MAS and 3QMAS NMR experiments on

kaolinite, a 1:1 layer silicate, and muscovite. These samples have not been

synthesised directly but were obtained from 17O exchange in natural samples during

4.6. 17O MAS and 17O 3Q-MAS NMR spectroscopy

137

0 -20 -40 ppm20406080

H-O-(Mg,Al)

[6] [6]Al-O- Al

Si-O- (Mg,Al)[6]

Si-O-Si 1

Si-O-Si 2

Si-O- Al[4]

whole fit

experiment

Figure 4.47. 17O MAS NMR spectrum of phlogopite with composition xnom = 0.5, y=1.0. Observed spectrum, total

lineshape fit, and individual signal components are shown.

hydrothermal synthesis. Not all of the oxygen positions show the same exchange

rate, thus the results are not directly comparable to the phlogopites where 17O has

been incorporated during crystallisation. Nevertheless, the signal assignments

proposed by these authors should be able to give some indications for the phlogopite

spectra.

Signals resulting from Si-O-[4]Al and Si-O-Si environments show a very similar

position in the F2-dimension, but they may be distinguished by their different δ(F1)

values. The two signals at δ(F2) 47 ppm, δ(F1) ~ 30 ppm and δ(F2) ~ 34 ppm,

δ(F1) ~ 35 ppm, respectively, may result from two Si-O-Si environments,

characterised by different T-O-T bond angles. Lee et al. (2003a) already observed a

splitting of signals for basal oxygen atom O4 on the one hand and for the basal

oxygen atoms O3 and O5 on the other hand in their 17O 3QMAS NMR spectra of

kaolinite. For both signals CQ is in the range of 2.6 to 3.0 MHz. Lee et al. (2003a)

found much larger quadrupolar coupling constants of 4.4 to 4.8 MHz. It should be

noted that the δ(F1)-values observed in this study are also very different from what

4. Results and discussion

138

Table 4.9. Quadrupolar coupling parameters obtained from fits of 17O MAS and MQMAS NMR spectra obtained at

9.34 T. Error ranges have been estimated by changing the parameters manually observing χ2 until a disting

change of χ2 took place.

signal δ(F2) CQ η δ(F1) F

component [ppm] ± 0.5 [MHz] ± 0.1 [ppm] ± 3 [%] ± 3

H-O-[6](Mg,Al) 78.6 5.5 ± 0.5 0.6 - 20

Si-O-[6](Mg,Al) 71.1 1.5 ± 0.3 0.7 ~ 37 6

Si-O-Si 1 46.9 2.8 ± 0.2 0.3 ~ 30 13

Si-O-[4]Al 42.5 2.2 ± 0.2 0.7 ~ 25 40

Si-O-Si 2 33.8 2.7 ± 0.2 0.3 ~ 35 5

Al-O-Al 25.0 2.4 ± 0.2 0.7 ~ 17 16

these authors reported. However, this is due to different labelling conventions of the

isotropic F1-dimension.

The Si-O-[4]Al signal is positioned at δ(F2) = 43 ppm, δ(F1) ~ 25 ppm. In

agreement with the findings of Lee et al. (2003a), its quadrupolar coupling constant

of 2.0 – 2.4 MHz is slightly smaller than that of the Si-O-Si signals due to differences

in the bond angles.

The position of the [4](Si,Al)-O-[6]Mg,Al) signal is expected to be shifted to less

shielded values compared to the Si-O-Si and Si-O-Al signals. The narrow signal at

δ(F2) = 71 ppm, δ(F1) ~ 37 ppm might result from this type of environment. The

quadrupolar coupling constant CQ of 1.5 MHz is again much smaller than that

observed by Lee et al. (2003a) who reported a CQ of about 3.5 MHz for apical oxygen

atoms.

The fifth signal observed in the 17O MQMAS NMR spectra has a position of

δ(F2) = 25 ppm, δ(F1) ~ 17 ppm. This is close to the position where a [4]Al-O-[4]Al

signal could be expected (Stebbins et al., 1999), and indeed its quadrupolar coupling

constant is similar to that of the Si-O-[4]Al signal (CQ = 2.4 ± 0.2 MHz). However, this

signal shows a strong electric field gradient distribution resulting from a variation of

the oxygen environment in the structure. Therefore it is more likely that the signal

results from [6]Al-O-[6]Al environments in the structure of aluminium oxide also present

in the sample, but the presence of [4]Al-O-[4]Al linkages in the phlogopite structure

cannot be ruled out completely.

4.7. Analysis of X-ray diffraction powder patterns

139

4.7. Analysis of X-ray diffraction powder patterns

17 representative phlogopite samples have been investigated using powder X-ray

diffraction (XRD) techniques. The resulting patterns can be used to help identifying

impurity phases in addition to the NMR spectroscopic experiments. Moreover, they

allow a distinction between different phlogopite polytypes formed during synthesis

(see Chapter. 2.1.3). Because possible ordering patterns are different for the three

polytypes (see Chapters 2.1.4 and 2.1.5), this information is vital in the investigation

of ordering in the phlogopite structure. The lattice parameters of the polytypes may

also be extracted from the data. Structural changes upon replacement of atoms by

differently sized atoms should also be become visible in a changing of the lattice

parameters. In this way, powder X-ray diffraction patterns can complement the NMR

spectroscopic investigations by allowing to view on the structure on a larger scale.

Figure 4.48 shows four selected diagrams. As expected, all samples show the

typical Bragg peaks for the 1M-polytype. However, many samples also exhibit

additional reflections resulting from the less common 2M1-polytype which differs from

x = 0.0y = 0.5

x = 0.7y = 0.5

x = 0.5y = 2.0

x = 1.6y = 2.0

Figure 4.48. X-ray diffraction powder patterns of four selected phlogopite samples. Arrows mark peak positions of

the impurity phase corundum.

4. Results and discussion

140

the 1M-polytype by a slightly smaller monoclinic angle and a doubling of the lattice

parameter c. Other reflections are due to corundum (α-Al2O3), and the increase of

this impurity phase with higher Al-contents of the initial gel composition can be

observed from a comparison of the patterns.

Other peaks (not shown) result from minor amounts of the impurity phases

potassium aluminium hexafluoride (K3AlF6*0.5H2O), sellaite (MgF2) or some chlorite-

type structure.

The phlogopite reflections of all samples show low intensity and broad half-widths

due to the low crystallinity of the samples. The crystallite sizes determined by LeBail-

fitting of the patterns with the Topas software (Bruker) are in the range between 40

and 120 nm (Table 4.10), and thus, slightly lower than those observed in the SEM

investigations (Chapter 4.1).

Table 4.10. Crystallite sizes and relative amounts of phases in phlogopite samples determined by LeBail-fitting of

the phlogopite XRD patterns. The given R-value is that of the overall fit.

x y crystallite

size [nm]a)

R-value

Phl-1M [%]b)

Phl-2M1 [%]b) Phl-1Mrel

c)

0.8 0.0 77 (4) 6.12 51 19 73

0.1 0.2 47 (2) 6.44 68 31 69

0.0 0.5 138 (6) 6.22 89 14 87

0.2 0.5 122 (5) 5.99 91 8 92

0.5 0.5 73 (4) 5.97 79 19 80

0.7 0.5 62 (3) 6.68 71 15 83

0.0 1.0 80 (5) 6.24 67 33 67

0.5 1.0 30 (1) 6.29 79 21 79

0.6 1.0 46 (3) 5.99 65 26 71

0.8 1.0 46 (3) 6.10 75 25 75

0.7 1.2 63 (3) 5.96 74 19 80

0.2 1.6 41 (2) 6.12 60 31 66

0.4 1.6 43 (2) 6.02 73 27 73

0.5 1.6 42 (2) 5.86 79 16 83

1.2 1.6 55 (2) 6.02 71 11 86

0.5 2.0 43 (2) 6.07 78 19 81

1.6 2.0 57 (2) 6.07 56 11 83

a) Errors are given in parenthesis.

b) The error bar of the relative amounts is ± 5 %.

c) Phl-1Mrel= Phl-1M / (Phl-1M + Phl-2M1)

4.7. Analysis of X-ray diffraction powder patterns

141

a [A

]

b [A

]

c [A

]

]

y = 0.0 y = 0.2 y = 0.5 y = 1.0

y = 1.2 y = 1.6 y = 2.0 Figure 4.49. Results of the analysis of XRD powder patterns of several 1M-phlogopites with varying compositions

K(Mg3-xAlx)(Al1+xSi3-xO10)(OH)yF2-y.

These features make it difficult to analyze the patterns, but still the lattice

parameters could be determined for the 1M polytype (Figure 4.49). A decrease in the

a and b lattice parameters with increasing incorporation of Al has been found: The

lattice parameter a is in the range of 5.315±0.005 Å at low Al-contents and

5.285±0.005 Å at high Al-contents, while the lattice parameter b ranges between

9.200±0.005 Å at low Al-contents and 9.155±0.005 Å at high Al-contents. This can be

explained by the replacement of Mg2+ with the smaller Al3+ cation. The substitution

leads to a shrinking of the octahedral sheets, at the same time increasing the lateral

misfit between octahedral and tetrahedral sheets. For compensation, a rotation of

tetrahedra is necessary, lowering the symmetry of the hexagonal rings from

hexagonal to ditrigonal (Figure 4.50). As a result the lateral dimensions of the whole

structure decrease within the a,b-plane.

4. Results and discussion

142

a b c

Figure 4.50. Sketch showing the distortion of a tetrahedral sheet. a) Undisturbed sheet with hexagonal symmetry.

b) Rotation of tetrahedra about the perpendicular to the sheet leads to a ditrigonal symmetry. The distortion is

described by the ditrigonal rotation angle α. c) Fully distorted tetrahedral sheet. (Ferraris and Ivaldi, 2002, p.131)

There has also been observed a variation in the lattice parameter c as has been

reported in the literature: For OH-rich samples this parameter is about 0.1 to 0.15 Å

higher than for F-rich samples. In OH-rich samples, the proton directly points towards

the interlayer cation K+ leading to a strong repulsion and a widening of the distance

between adjacent layer packages. With F substituting for OH this repulsion is

reduced leading to a narrowing of the interlayer area (for details see Chapter 2.1.6).

The lattice parameter c also seems to increase with increasing Al-content. However,

this could also be due to the fact that no high-Al samples have been prepared for

lower F-contents. For given F-content there is no significant increase in the lattice

parameter with higher Al-contents. No correlation can be identified for the monoclinic

angle β.

For the 2M1-polytype, the data points scatter much more and no correlations

have been observed (Figure 4.51). This can be explained by the lower amount of this

polytype in the mixture compared to phlogopite-1M (Table 4.9) resulting in lower

quality data.

Phlogopite-1M dominates the synthesised mixture of phlogopite polytypes. The

relative amount of this polytype determined from the LeBail-fits ranges from 65 to

93 % (Figure 4.52). No correlation between the relative amounts of both polytypes

and the composition of the phlogopites has been observed, which is in agreement

with Bigi et al. (1993) who reported the same for natural samples containing all three

phlogopite polytypes with a high degree of disorder and stacking faults. Nevertheless

4.7. Analysis of X-ray diffraction powder patterns

143

c [A

]

b [A

]

a [A

]

]

y = 0.0 y = 0.2 y = 0.5 y = 1.0

y = 1.2 y = 1.6 y = 2.0 Figure 4.51. Results of the analysis of XRD powder patterns of several 2M1-phlogopites with varying compositions

K(Mg3-xAlx)(Al1+xSi3-xO10)(OH)yF2-y.

Figure 4.52. The relative amount of phlogopite-1M of phlogopites derived from LeBail-fitting plotted against the

estimated Al-content of the phlogopites.

4. Results and discussion

144

the appearance of polytype 2M1 is best explained by structural changes upon

incorporation of Al. Higher amounts of Al in the octahedral sheets lead to a higher

ditrigonal rotation in tetrahedral sheets, as has been mentioned before. However,

another effect is the tendency for a more ‘dioctahedral character’ of Al-rich

phlogopites. Due to the different cationic radii of Mg2+ and Al3+ the difference in size

between separate octahedra is increased. This in return leads to a tilting of

tetrahedra out of the (001) plane (see review by Ferraris and Ivaldi, 2002, and

references therein). In polytype 1M, the tetrahedra of opposing layer packages at the

interlayer boundary show a tilt in opposite directions, i.e. away from each other

(Figure 4.53). This leads to a widening of the interlayer cation site. In phlogopite-2M1

however the tetrahedra are tilting in the same direction and the size of the interlayer

cation site does not change. Therefore, exchange of Mg2+ ↔ Al3+ leads to a

destabilisation of the 1M polytype, and the formation of phlogopite-2M1 is forced

instead.

Figure 4.53. Sketch of the interlayer boundary in phlogopite-1M (left) and phlogopite 2M1 (right). Tetrahedral

tilting, i.e. out-of-plane rotation, is exaggerated. Circles denote K+-ions. Modified after Ferraris and Ivaldi, 2002, p.

134.

The presence of two polytypes leads to a significant number of stacking faults

and disorder in the structure. As a result the observed powder XRD patterns all show

a high background in form of broad bumps (Figure 4.48). Moreover, satellite

reflections have been observed in more detailed measurements (Figure 4.54). These

reflections did not disappear upon heating up to 500 °C, indicating that they do not

result from a modulation but from structural disorder. Stacking faults increase the

periodicity of the lattice and lead to a larger supercell, owing to weak superstructure

reflections.

4.7. Analysis of X-ray diffraction powder patterns

145

Figure 4.54. XRD pattern of phlogopite with nominal composition xnom = 0.4, y = 1.8. a) Whole pattern. b) Detail.

The pattern shows a high background, and the peaks between 20 and 33 °2θ are surrounded by satellite peaks

(marked by arrows) resulting from stacking faults in the structure.

For comparison, phlogopite samples synthesised at 800 °C and described earlier

by Fechtelkord et al. (2003 a,b) have been analysed, too. Data have been measured

by Dr. Karen Friese and Dr. Andrzej Grzechnik at beamline D3 at HASYLAB, DESY,

Hamburg, Germany, and friendly provided for investigation in the frame of this study.

LeBail-fitting of the observed patterns was performed in collaboration with Dr. Karen

Friese and Dr. Andrzej Grzechnik at the Universidad del Pais Vasco, Bilbao, Spain.

Three fitting programmes have been used independently: the JANA2000 software

(Dušek et al., 2001), the UnitCell software by Tim Holland and Simon Redfern

(Department of Earth Sciences, University of Cambridge, UK), and the Chekcell

powder indexing tool by by Jean Laugier and Bernard Bochu (Laboratoire des

Materiaux et du Génie Physique de l'Ecole Supérieure de Physique de Grenoble,

France). The results were similar and thus, only the results obtained from Jana are

shown in Figure 4.55.

4. Results and discussion

146

a [A

]

b [A

]

c [A

]

]

y = 1.0 y = 1.5y = 0.5 Figure 4.55. Results of the analysis of XRD powder patterns of several 2M1-phlogopites with varying compositions

K(Mg3-xAlx)(Al1+xSi3-xO10)(OH)yF2-y.

As has already been observed for the samples synthesised at 600 °C, the lattice

parameter a decreases with increasing Al-incorporation, and the lattice parameter c

increases with OH-content of the phlogopites. The lattice parameter b is more or less

constant now, in agreement with the literature. Nevertheless, these results show that

the same structural changes take place on incorporation of Al and F both at 600 and

at 800 °C.

5. Conclusions and outlook

147

5. Conclusions and outlook

All of the experimental and computational results have shown that incorporation

of Al into the phlogopite structure is energetically unfavourable. For all samples of

significant Al-contents the amount of Al in the phlogopite structure has been found to

be lower than that of the initial gel composition, and Al-rich impurity phases – mostly

Al2O3 – have been formed.

The results indicate that the ability to incorporate Al does not differ much for

phlogopites of varying F-contents. However, large differences have been observed

for extreme conditions, i.e. F-free and OH-free oxide mixtures. While hydroxyl-

phlogopites always showed a higher Al-contents than F-containing compositions, the

amount of Al incorporated into the water-free phlogopite was much lower.

A saturation effect has been found: The highest Al-content of pure hydroxyl-

phlogopites has been reached at xest = 0.82, corresponding to a composition of

K (Mg2.17Al0.82) (Al1.82Si2.17O10) (OH)2, and further addition of Al to the initial oxide

mixture did not yield more Al-rich phlogopites. A comparison between samples

synthesised at 600 and at 800 °C did not show significant differences, although a

positive effect on the incorporation of Al could have been expected for lower

temperatures.

Even if the amount of F in the initial oxide mixture did not change the Al-content

of the synthesised phlogopites, it had a strong influence on the amount of impurity

phases formed during synthesis. It has been found that high amounts of F prevented

extensive formation of Al-rich phlogopites and vice versa, resulting in a formation of

K3AlF6*0.5H2O instead of phlogopite (in addition to Al2O3). This effect has not only

been observed for F-rich samples and higher Al-contents, but also for extremely Al-

rich compositions even if the amount of F was very low.

For hydroxyl-phlogopites the Al-content of the octahedral sheet could be

determined and compared to that of the tetrahedral sheet. For Al-rich phlogopites

both values agreed well, but for Al-poor phlogopites the amount of Al estimated for

the octahedral sheet was higher than that of the tetrahedral sheet. This could result

from other substitution mechanism than Tschermak’s substitution taking place at low

Al-contents. Another possible explanation is the presence of Al-O-Al linkages in the

tetrahedral sheets resulting in an underestimation of the real amount of [4]Al.

5. Conclusions and outlook

148

The disfavour of phlogopite to incorporate additional Al into its structure is also

visible on the atomic level. The 29Si MAS NMR spectra indicate complete solid-

solution between phlogopite and ‘eastonite’, but this is only true on the macroscopic

level. Further NMR spectroscopic experiments as well as Monte Carlo simulations of

cation ordering in hydroxyl-phlogopites showed that in fact, the ‘eastonite’ component

is incorporated into the phlogopite structure in form of clusters affecting all sheets of

a single layer package. At x = 0.0 only K Mg3 (AlSi3O10) (OH)2 is present. As soon as

additional Al is brought into the structure, small clusters of composition

K (Mg2Al) (Al2Si2O10) (OH)2 appear. On further increase of the Al-content, these

clusters are enlarged, until – hypothetically – at x = 1.0 all areas of phlogopite

composition have disappeared.

The strain imposed on the structure on replacement of Mg/Si by [6]Al/[4]Al leads to

the formation of disordered phlogopites, composed of two different polytypes. Most of

the mixture is made up by polytype 1M which is typical for trioctahedral micas.

However, structural changes due to the substitution of Mg by the smaller Al lead to

the formation of phlogopite-2M1 which has a more dioctahedral character with

differently sized octahedra. A high background in the XRD patterns and satellite

reflections indicate a high degree of disorder, resulting from intergrowth of both

polytypes.

The observed ordering pattern can be traced back to three different interactions

controlling the distribution of ions in the phlogopite structure. In tetrahedral sheets,

ordering is dominated by the avoidance of Al-atoms as next-nearest-neighbours,

according to Loewenstein’s rule. This induces perfect ordering in the Al-rich clusters,

where Si and [4]Al strictly occupy tetrahedral sites alternately. The same is true for the

octahedral sheet, where Al is always surrounded by six Mg-ions. Again, placing Al-

atoms on directly neighboured sites is highly unfavourable. In contrast, there is a

strong preference for Al to occupy adjacent octahedral and tetrahedral sites, leading

to the cluster formation mentioned before. This means, Al-O-Al linkages are indeed

favourable in the phlogopite structure, if two different types of polyhedra are involved.

The strong preference of OH and F to be co-ordinated by Al-rich and Mg-rich

environments, respectively, suggests enrichment of OH in the ‘eastonite’ clusters. In

NMR spectroscopic experiments, a grouping of H and F in the octahedral sheet has

been observed. Moreover, it has been shown that the Al-content of the tetrahedral

5. Conclusions and outlook

149

sheet close to F is slightly lower than the overall Al-content. This indicates that a

separation into K Mg3 (AlSi3O10) F2 and K (Mg2Al) (Al2Si2O10) (OH)2 clusters is

favoured. However, this is not always possible depending on the F- and OH-contents

of the phlogopites. 1H and 19F MAS NMR spectra clearly show a variation in the

Al/Mg contents of OH/F environments with increasing amounts of Al in the structure.

It has been shown that the combination of NMR spectroscopy and MC

simulations is a useful tool for the investigation of ordering mechanisms in

phlogopite. A relationship between the ordering of ions in both sheets has been

clearly identified, and it has been demonstrated that the observed clustering has

important crystal chemical consequences influencing the overall phlogopite structure

and composition.

However, further investigations are still necessary. OH/F ordering has not been

considered in the simulations yet, but additional information is necessary to clarify the

observed relationship between ‘eastonite’ clustering and OH/F distribution.

Further investigation is also necessary from the experimental point of view.

Additional synthesis runs at T = 400 °C are necessary to improve our understanding

of the influence of temperature on Al-incorporation. Moreover, the chemical

composition of the investigated phlogopites has been very limited. The results

obtained in this study form a valuable basis, but compositions closer to natural

phlogopites/biotites need to be studied, too. Addition of small amounts of iron to the

starting mixture might be possible without resulting in too much broadening of NMR

lineshapes.

The methods used here may also be transferred to other mica structures.

Computational studies are already available for the dioctahedral muscovite/phengite

series (Palin and Dove, 2004) showing a Mg/[4]Al preference that is contradictory to

what was found for phlogopite in this study. NMR spectroscopic investigations are

necessary to obtain further information on cation ordering in phengite and to clarify

the fundamental differences between ordering mechanism in both types of structure.

A. Appendix

150

A. Appendix

A.1. List of abbreviations

Table 6.1. List of abbreviations

symbol / abbreviation explanation

[4]Al Aluminium in tetrahedral co-ordination [6]Al Aluminium in octahedral co-ordination

a Lattice parameter a of the unit cell

at% Atomic percentage

b Lattice parameter b of the unit cell

B0 External magnetic field

β Monoclinic angle

c Lattice parameter c of the unit cell

CQ Quadrupolar coupling constant

CPMAS Cross-polarisation magic angle spinning

Crn Corundum

CS Chemical shift

E Energy

EDX Energy-dispersive X-ray detector system

EFG Electric field gradient

EMP Electron microprobe

F1, F2 Dimensions in the MQMAS experiment

h Planck’s constant

H Hamiltonian of first-order interactions

CSH Hamiltonian of chemical shift interaction

A.1. List of abbreviations

151

symbol / abbreviation explanation

DDH Hamiltonian of dipolar interaction

QH Hamiltonian of quadrupolar interaction

zH Hamiltonian of Zeeman interaction

HETCOR CPMAS Hetero-nuclear correlation cross-polarisation magic angle spinning

I Nuclear spin of the observed nucleus

IR spectroscopy Infra-red spectroscopy

kB Boltzmann-constant

KAF Potassium aluminium hexafluoride (K3AlF6*0.5H2O)

LS Laboratory axes system

µ Magnetic moment

µ0 Permeability of vacuum

m Magnetic quantum number

M Molarity

M1 Octahedral site in the phlogopite structure

M2 Octahedral site in the phlogopite structure

MAS Magic angle spinning

MQMAS Multiple-quantum magic angle spinning

Msc Muscovite (K Al2 (AlSi3O10) (OH,F)2)

η Asymmetry parameter

NMR spectroscopy Nuclear magnetic resonance spectroscopy

p Pressure

PAS Principle axes system

Phl Phlogopite (K Mg3 (AlSi3O10) (OH,F)2

phlogopite-1M Phlogopite of the 1M-polytype

phlogopite-2M1 Phlogopite of the 2M1-polytype

A. Appendix

152

symbol / abbreviation explanation

phlogopite-2O Phlogopite of the 2O-polytype

phlogopite-3T Phlogopite of the 3T-polytype

Q Quadrupolar moment

σ Chemical shielding tensor

σiso Isotropic chemical shift

σaniso Chemical shift anisotropy

σQS Quadrupolar shift

S Nuclear spin of a nucleus not under observation

SE Secondary electrons

SEM Scanning electron microscopy

T Temperature

THSi Cross-polarisation time

T1ρ Spin-lattice relaxation time

TEOS Tetraethyl orthosilicate

TMS Tetramethylsilane

0 Larmor frequency

Q Quadrupolar frequency

rf Frequency of the radio-frequency pulse

ω0 Larmor precession frequency

ωref Resonance precession frequency of the observed nucleus

ωx Resocance precession frequency of the reference material

wt% Weight percentage

xest Estimated real Al-content of the synthesised phlogopites

xnom Nominal Al-content of the initial gel composition K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y

XRD X-ray diffraction

γ Magnetogyric ratio

A.2. NMR spectroscopic results

153

A.2. NMR spectroscopic results

Tab

le 6

.2. P

aram

eter

s ob

tain

ed fr

om 2

9S

i MA

S N

MR

spe

ctra

.

“cla

y”5)

A4)

[%]

2 3 2 3 7

FW

HM

3

) [p

pm]

2.9

2.9

2.2

2.1

2.9

pos.

2)

[ppm

]

-95.

7

-94.

8

-94.

7

-94.

8

-94.

7

Si-O

-Si 3

A4)

[%]

7 20

19

22

14

16

7 12

9 6 4 14

6 6 7 20

15

17

FW

HM

3

) [p

pm]

2.4

2.8

2.0

2.8

2.6

2.8

2.8

2.8

2.8

2.8

2.8

2.8

2.8

2.6

2.8

2.5

2.8

2.8

pos.

2)

[ppm

]

-90.

8

-93.

1

-92.

4

-91.

9

-91.

8

-91.

8

-90.

5

-91.

9

-90.

6

-91.

3

-91.

2

-91.

9

-91.

4

-90.

6

-91.

0

-92.

7

-91.

7

-91.

4

Si-O

-AlS

i 2

A4)

[%]

42

58

62

57

60

54

47

45

41

44

41

57

40

37

36

60

55

50

FW

HM

3

) [p

pm]

2.3

2.4

1.9

2.2

2.5

2.7

2.5

2.5

2.7

2.6

2.6

2.4

2.5

2.4

2.5

2.1

2.4

2.4

pos.

2)

[ppm

]

-87.

5

-88.

9

-88.

4

-87.

9

-88.

2

-88.

2

-86.

9

-88.

0

-86.

9

-87.

8

-87.

5

-88.

3

-87.

4

-87.

1

-87.

4

-88.

6

-87.

9

-88.

0

Si-O

-Al 2

Si

A4)

[%]

43

22

20

21

26

27

42

36

39

44

45

29

45

47

44

19

29

26

FW

HM

3

) [p

pm]

2.4

2.5

2.2

2.6

2.6

2.7

2.6

2.7

2.6

2.6

2.6

2.6

2.6

2.9

2.6

2.2

2.6

2.6

pos.

2)

[ppm

]

-84.

1

-85.

1

-84.

7

-84.

2

-84.

7

-84.

2

-83.

3

-84.

3

-83.

5

-84.

3

-84.

1

-84.

7

-83.

8

-83.

5

-83.

8

-84.

7

-84.

1

-84.

3

Si-O

-Al 3

A4)

[%]

7 4 7 10

6 11 8 11

13

FW

HM

3

) [p

pm]

2.9

2.8

2.8

2.8

2.8

2.8

2.8

2.8

2.8

pos.

2)

[ppm

]

-80.

4

-80.

0

-80.

7

-80.

7

-80.

8

-80.

9

-80.

2

-79.

6

-80.

2

x est

1)

0.24

0.01

0.01

0.00

0.09

0.08

0.29

0.26

0.33

0.33

0.40

0.11

0.37

0.40

0.41

0.00

0.10

0.07

y

0.0

0.2

0.5

0.5

0.5

0.5

0.5

0.5

0.5

0.5

0.5

0.8

0.8

0.8

0.8

1.0

1.0

1.0

x

0.8

0.1

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.2

0.4

0.6

0.8

0.0

0.1

0.2

A. Appendix

154

T

able

6.2

. Co

ntin

ued.

“cla

y”5)

A4)

[%] 6 3 4 5 6 4 7 9 7 8

FW

HM

3)

[ppm

]

2.9

2.9

2.1

2.9

2.9

2.9

2.9

2.9

2.9

2.9

pos.

2)

[ppm

]

-95.

8

-95.

2

-95.

2

-95.

5

-95.

3

-95.

2

-95.

8

-95.

1

-93.

6

-93.

8

Si-O

-Si 3

A4)

[%]

11

17

14

7 7 7 19

16

16

17

15

15

7 19

23

14

16

15

FW

HM

3)

[ppm

]

2.8

2.8

2.8

2.8

2.8

2.8

2.8

2.9

2.8

2.8

2.8

2.8

2.1

2.1

2.8

2.8

2.8

2.9

pos.

2)

[ppm

]

-91.

4

-91.

7

-91.

9

-90.

3

-90.

7

-90.

4

-92.

3

-92.

5

-91.

4

-91.

3

-91.

3

-91.

3

-91.

3

-92.

4

-91.

2

-90.

8

-91.

1

-92.

3

Si-O

-AlS

i 2

A4)

[%]

49

45

47

36

27

28

56

53

44

45

42

40

43

63

49

46

52

53

FW

HM

3)

[ppm

]

2.4

2.7

2.6

2.7

2.5

2.7

2.6

2.4

2.7

2.7

2.7

2.7

2.6

2.0

2.4

2.7

2.7

2.5

pos.

2)

[ppm

]

-87.

7

-87.

9

-88.

1

-86.

8

-87.

2

-87.

0

-88.

2

-88.

5

-87.

7

-87.

8

-87.

7

-87.

7

-87.

6

-88.

4

-87.

7

-87.

8

-87.

8

-88.

5

Si-O

-Al 2

Si

A4)

[%]

36

26

29

43

43

42

25

31

29

29

30

30

40

18

20

28

25

33

FW

HM

3)

[ppm

]

2.6

2.6

2.7

2.5

2.5

2.6

2.8

2.9

2.6

2.8

2.6

2.7

2.6

2.1

2.6

2.7

2.5

2.9

pos.

[ppm

]

-84.

0

-84.

0

-84.

3

-83.

2

-83.

6

-83.

4

-84.

1

-84.

7

-83.

9

-83.

9

-83.

8

-83.

8

-83.

9

-84.

4

-84.

0

-84.

1

-84.

0

-84.

5

Si-O

-Al 3

A4)

[%]

4 6 6 13

23

23 5 3 9 8 10 4

FW

HM

3)

[ppm

]

2.8

2.8

2.8

2.7

2.6

2.7

2.8

2.7

2.8

2.8

2.6

2.8

pos.

2)

[ppm

]

-80.

3

-80.

5

-80.

5

-79.

5

-79.

9

-79.

7

-80.

4

-80.

0

-79.

9

-79.

9

-80.

1

-80.

5

x est

1)

0.23

0.16

0.20

0.41

0.51

0.51

0.04

0.11

0.18

0.14

0.24

0.23

0.35

0.00

0.00

0.17

0.07

0.13

y

1.0

1.0

1.0

1.0

1.0

1.0

1.2

1.2

1.2

1.2

1.2

1.2

1.2

1.4

1.4

1.4

1.4

1.4

x

0.3

0.4

0.5

0.6

0.7

0.8

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.0

0.1

0.2

0.3

0.4

A.2. NMR spectroscopic results

155

T

able

6.2

. Co

ntin

ued.

“cla

y”5)

A4)

[%]

6 9 5 5 10

2 6 4

FW

HM

3)

[ppm

]

2.9

2.9

2.9

2.9

2.9

3.1

2.9

2.9

pos.

2)

[ppm

]

-94.

3

-95.

2

-94.

3

-94.

4

-93.

8

-95.

5

-93.

2

-94.

3

Si-O

-Si 3

A4)

[%]

14

10

7 22

20

17

21

11

10

10

6 6 14

17

13

5 6 7

FW

HM

3)

[ppm

]

2.8

2.8

2.8

2.8

2.8

2.8

2.8

2.8

2.8

2.8

2.9

2.9

2.8

2.8

2.8

2.1

2.9

2.9

pos.

2)

[ppm

]

-91.

6

-91.

4

-91.

9

-91.

0

-91.

2

-91.

4

-90.

2

-90.

7

-91.

2

-91.

3

-91.

1

-90.

7

-90.

3

-90.

3

-90.

5

-90.

2

-90.

2

-89.

6

Si-O

-AlS

i 2

A4)

[%]

49

45

44

46

52

51

40

38

47

38

21

16

50

44

44

35

14

14

FW

HM

3)

[ppm

]

2.7

2.6

2.7

2.6

2.7

2.7

2.7

2.7

2.5

2.7

2.8

2.8

2.8

2.6

2.7

2.4

2.8

2.8

pos.

2)

[ppm

]

-88.

1

-87.

6

-87.

6

-87.

4

-87.

6

-87.

8

-87.

4

-87.

0

-87.

3

-87.

4

-87.

2

-87.

3

-86.

8

-86.

8

-87.

3

-86.

6

-86.

7

-86.

1

Si-O

-Al 2

Si

A4)

[%]

27

37

37

20

24

27

25

38

35

40

34

38

27

33

35

43

30

36

FW

HM

3)

[ppm

]

2.7

2.7

2.6

2.7

2.6

2.6

2.7

2.6

2.3

2.7

2.4

2.6

2.6

2.4

2.6

2.4

2.4

2.5

pos.

2)

[ppm

]

-84.

1

-83.

9

-83.

9

-83.

5

-83.

6

-84.

0

-83.

7

-83.

3

-83.

5

-83.

7

-83.

7

-83.

6

-82.

8

-83.

2

-83.

5

-83.

0

-83.

0

-82.

8

Si-O

-Al 3

A4)

[%]

5 7 12

3 4 13

8 11

39

40

3 4 8 17

50

44

FW

HM

3)

[ppm

]

2.8

2.8

2.8

2.8

2.8

2.9

2.8

2.8

1.9

1.8

2.8

2.7

2.8

2.3

1.7

1.8

pos.

2)

[ppm

]

-80.

4

-80.

1

-80.

1

-79.

6

-80.

0

-79.

6

-79.

8

-80.

1

-80.

1

-80.

0

-79.

0

-79.

7

-80.

3

-79.

5

-79.

4

-79.

4

x est

1)

0.17

0.29

0.36

0.03

0.03

0.08

0.1

0.4

0.3

0.4

0.6

0.7

0.2

0.2

0.3

0.5

0.7

0.7

y

1.4

1.4

1.4

1.6

1.6

1.6

1.6

1.6

1.6

1.6

1.6

1.6

1.8

1.8

1.8

1.8

1.8

1.8

x

0.5

0.6

0.7

0.1

0.2

0.3

0.4

0.5

0.6

0.8

1.0

1.2

0.2

0.3

0.5

0.6

1.2

1.6

A. Appendix

156

T

able

6.2

. Co

ntin

ued.

“cla

y”5)

A4)

[%]

2 6

1)

x

est =

Add

ition

al A

l-con

tent

of p

hlo

gop

ites

dete

rmin

ed f

rom

29S

i MA

S N

MR

spe

ctra

. Err

or r

ang

e is

± 0

.1.

2)

p

os. =

Sig

nal p

ositi

on.

Err

or r

ang

e is

± 0

.3 p

pm.

3)

F

WH

M =

Ful

l wid

th a

t ha

lf m

axim

um. E

rror

ran

ge

is ±

0.4

ppm

.

4)

A

= R

elat

ive

sign

al a

rea.

Err

or r

ang

e is

± 3

%.

5)

“C

lay”

des

crib

es a

K-d

efic

ient

bru

cite

like

laye

r in

the

stru

ctur

e.

FW

HM

3)

[ppm

]

2.9

2.8

pos.

2)

[ppm

]

-93.

8

-94.

7

Si-O

-Si 3

A4)

[%]

13

14

5 3

FW

HM

3)

[ppm

]

2.8

2.9

2.4

2.9

pos.

2)

[ppm

]

-90.

8

-90.

9

-89.

8

-90.

0

Si-O

-AlS

i 2

A4)

[%]

43

37

17

8 8 19

FW

HM

3)

[ppm

]

2.7

2.9

2.7 2 2.6

2.7

pos.

2)

[ppm

]

-87.

6

-87.

1

-86.

3

-86.

2

-86.

7

-86.

6

Si-O

-Al 2

Si

A4)

[%]

34

34

36

31

20

42

FW

HM

3)

[ppm

]

2.6

2.7

2.5

2.5

2.6

2.6

pos.

2)

[ppm

]

-83.

7

-83.

1

-82.

8

-82.

6

-82.

8

-82.

9

Si-O

-Al 3

A4)

[%]

7 10

42

61

72

36

FW

HM

3)

[ppm

]

2.8

2.9

2.3

1.4

1.3

2.0

pos.

2)

[ppm

]

-80.

1

-79.

8

-79.

4

-79.

4

-79.

6

-79.

4

x est

1)

0.3

0.3

0.7

0.8

0.8

0.7

y

2.0

2.0

2.0

2.0

2.0

2.0

x

0.4

0.5

0.8

1.0

1.2

1.6

A.2. NMR spectroscopic results

157

Table 6.3. NMR spectroscopic parameters obtained from 1H MAS NMR spectra.

H-OMg3 H-OMg2Al

x y xest1)

pos.2)

[ppm]

FWHM3)

[ppm]

A4)

[%]

pos.1)

[ppm]

FWHM3)

[ppm]

A4)

[%]

OHAl 5

(OHAl+OHMg)

0.8 0.0 0.24 0.4 1.0 46 1.6 1.6 54 0.54

0.1 0.2 0.01 0.2 0.9 71 1.5 1.2 29 0.29

0.0 0.5 0.01 0.2 0.7 100 0.00

0.1 0.5 0.00 0.2 0.8 79 1.4 1.3 21 0.21

0.2 0.5 0.09 0.4 0.8 57 1.5 1.3 43 0.43

0.3 0.5 0.08 1.2 0.9 60 2.4 1.3 40 0.40

0.4 0.5 0.29 0.6 0.9 46 1.9 1.2 54 0.54

0.5 0.5 0.26 0.5 0.8 46 1.8 1.3 54 0.54

0.6 0.5 0.33 0.7 0.9 30 2.0 1.4 70 0.70

0.7 0.5 0.33 0.7 0.9 32 1.9 1.2 68 0.70

0.8 0.5 0.40 0.7 0.9 33 2.0 1.3 67 0.70

0.2 0.8 0.11 0.4 0.9 52 1.6 1.4 48 0.48

0.4 0.8 0.37 0.6 0.9 50 1.8 1.1 50 0.50

0.6 0.8 0.40 0.7 0.9 37 1.9 1.3 63 0.63

0.8 0.8 0.41 0.7 0.9 35 2.0 1.3 65 0.65

0.0 1.0 0.00 0.1 0.8 48 1.4 1.0 3 0.07

0.1 1.0 0.10 0.3 0.9 67 1.6 1.2 33 0.33

0.2 1.0 0.07 0.4 1.0 61 1.6 1.3 39 0.39

0.3 1.0 0.23 0.4 0.9 61 1.7 1.2 39 0.39

0.4 1.0 0.16 0.5 1.0 51 1.7 1.4 49 0.49

0.5 1.0 0.20 0.5 1.1 49 1.8 1.5 51 0.51

0.6 1.0 0.41 0.5 1.0 47 1.8 1.4 53 0.53

0.7 1.0 0.51 0.8 1.1 32 2.2 1.3 68 0.68

0.8 1.0 0.51 0.7 1.3 46 2.1 1.3 54 0.54

0.1 1.2 0.04 0.3 0.9 62 1.5 1.3 38 0.38

0.2 1.2 0.11 0.4 1.1 62 1.7 1.3 38 0.38

0.3 1.2 0.18 0.5 1.0 49 1.7 1.4 51 0.51

0.4 1.2 0.14 0.4 1.1 46 1.6 1.6 54 0.54

0.5 1.2 0.24 0.3 0.9 56 1.5 1.6 44 0.44

0.6 1.2 0.23 0.5 1.1 45 1.8 1.5 55 0.55

0.7 1.2 0.35 0.6 1.0 43 1.9 1.4 57 0.57

A. Appendix

158

Table 6.3. Continued.

H-OMg3 H-OMg2Al

x y xest1)

pos.2)

[ppm]

FWHM3)

[ppm]

A4)

[%]

pos.1)

[ppm]

FWHM3)

[ppm]

A4)

[%]

OHAl 5)

(OHAl+OHMg)

0.0 1.4 0.00 0.0 0.8 100 0.00

0.1 1.4 0.00 0.2 1.0 73 1.4 1.4 27 0.27

0.2 1.4 0.17 0.3 1.0 69 1.5 1.3 31 0.31

0.3 1.4 0.07 1.2 1.0 59 2.4 1.4 41 0.41

0.4 1.4 0.13 0.3 0.9 54 1.5 1.5 46 0.46

0.5 1.4 0.17 0.4 0.9 51 1.6 1.5 49 0.49

0.6 1.4 0.29 0.4 0.9 56 1.6 1.5 44 0.44

0.7 1.4 0.36 1.3 1.0 46 2.6 1.3 54 0.54

0.1 1.6 0.03 0.1 1.0 74 1.2 1.4 26 0.26

0.2 1.6 0.03 0.2 1.0 66 1.5 1.3 34 0.34

0.3 1.6 0.08 0.3 0.9 54 1.5 1.5 46 0.46

0.4 1.6 0.10 0.4 1.0 58 1.5 1.4 42 0.42

0.5 1.6 0.35 0.5 1.1 51 1.8 1.3 39 0.49

0.6 1.6 0.28 0.4 1.0 56 1.7 1.4 44 0.44

0.8 1.6 0.35 0.6 1.0 46 1.8 1.3 54 0.54

1.0 1.6 0.63 1.5 1.0 34 2.8 1.4 66 0.66

1.2 1.6 0.66 1.1 1.2 29 2.4 1.2 71 0.71

0.2 1.8 0.15 0.2 1.0 67 1.4 1.3 33 0.33

0.3 1.8 0.17 0.3 1.0 63 1.5 1.3 37 0.37

0.5 1.8 0.27 0.5 1.1 54 1.7 1.3 46 0.46

0.6 1.8 0.46 0.5 1.1 57 1.8 1.2 43 0.43

1.2 1.8 0.71 1.0 1.0 26 2.3 1.2 74 0.74

1.6 1.8 0.68 1.0 0.9 20 2.3 1.3 80 0.80

0.4 2.0 0.25 0.3 1.2 63 1.6 1.4 37 0.37

0.5 2.0 0.29 0.4 1.2 53 1.7 1.4 47 3.47

0.8 2.0 0.67 1.5 1.8 35 3.0 1.8 65 0.65

1.0 2.0 0.83 1.9 1.1 22 3.2 1.1 78 0.78

1.2 2.0 0.82 1.3 1.2 19 2.5 1.0 81 0.81

1.6 2.0 0.65 1.7 1.1 32 3.0 1.3 68 0.68 1. xest = Additional Al-content of phlogopites determined from 29Si MAS NMR spectra. Error range is ± 0.1.

2. pos. = Signal position. Error range: ± 0.1 ppm.

3. FWHM = Full width at half maximum. Error range: ± 0.2 ppm.

4. A = Relative signal area. Error range is ± 3 %.

5. OHAl/(OHAl+OHMg) = relative H-OMg2Al signal intensity. Error range: ± 0.05.

A.2. NMR spectroscopic results

159

Tab

le 6

.4. P

aram

eter

s ob

tain

ed fr

om 19

F M

AS

NM

R s

pect

ra.

othe

rs

A

[%]

27

6 1 2 3 13

5 7 9 28

8 23

2 20

4 1

K3A

lF6*

0.5H

2O A

[%] 6 1 10

FW

HM

[ppm

]

2.8

1.7

2.1

pos.

2)

[ppm

]

-158

.0

-157

.8

-158

.0

0.07

0.02

0.03

0.06

0.04

0.10

0.08

0.20

0.18

0.20

0.11

0.15

0.16

0.18

0.00

0.04

0.04

0.09

0.07

0.05

F-M

g 2A

l

A

[%]

5 2 3 6 3 10

7 17

17

15

9 10

14

14 3 3 9 7 5

FW

HM

[ppm

]

4.2

2.9

5.6

3.8

3.0

3.0

3.2

4.0

3.9

3.8

3.8

3.7

3.8

4.2

2.3

2.8

3.7

3.3

3.3

pos.

2)

[ppm

]

-150

.7

-151

.3

-151

.1

-151

.0

-150

.7

-150

.2

-151

.0

-150

.2

-150

.1

-150

.0

-150

.4

-149

.8

-149

.8

-149

.6

-150

.5

-151

.1

-150

.2

-150

.2

-150

.3

F-M

g 3

A

[%]

68

92

100

97

93

95

87

80

72

75

66

91

61

77

63

98

78

92

89

93

95

FW

HM

[ppm

]

3.2

2.8

2.4

2.6

2.8

2.9

3.4

3.3

3.8

3.7

3.6

3.0

3.4

3.7

3.8

2.4

3.0

3.0

3.1

3.3

3.1

pos.

2)

[ppm

]

-175

.3

-175

.9

-176

.1

-175

.4

-175

.4

-175

.3

-174

.1

-175

.2

-173

.8

-174

.2

-173

.7

-174

.7

-173

.5

-173

.8

-173

.6

-175

.8

-175

.1

-175

.5

-174

.4

-174

.9

-175

.3

x est

1)

0.24

0.01

0.01

0.00

0.09

0.08

0.29

0.26

0.33

0.33

0.40

0.11

0.37

0.40

0.41

0.00

0.10

0.07

0.23

0.16

0.20

y

0.0

0.2

0.5

0.5

0.5

0.5

0.5

0.5

0.5

0.5

0.5

0.8

0.8

0.8

0.8

1.0

1.0

1.0

1.0

1.0

1.0

x

0.8

0.1

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.2

0.4

0.6

0.8

0.0

0.1

0.2

0.3

0.4

0.5

A. Appendix

160

Tab

le 6

.4. C

on

tinue

d.

othe

rs

A4)

[%]

45

5 43

19

5

K3A

lF6*

0.5H

2O A

4)

[%]

7

FW

HM

3)

[ppm

]

2.7

pos.

2)

[ppm

]

-157

.9

FA

l 5)

(FA

l+F

Mg)

0.13

0.21

0.22

0.03

0.04

0.07

0.08

0.03

0.11

0.11

0.00

0.03

0.03

0.04

0.04

0.09

0.05

0.11

0.00

0.02

0.04

F-M

g 2A

l

A4)

[%]

7 18

13

3 3 7 8 3 11

11 3 3 4 4 9 5 11 2 4

FW

HM

3)

[ppm

]

4.1

4.4

5.4

2.6

3.3

3.1

3.6

2.8

4.2

3.5

2.9

3.2

2.6

3.1

3.7

3.6

3.6

2.5

2.6

pos.

2)

[ppm

]

-149

.3

-149

.1

-148

.8

-150

.5

-150

.7

-150

.3

-149

.8

-151

.3

-149

.9

-149

.8

-151

.0

-150

.5

-150

.4

-150

.9

-150

.0

-150

.1

-149

.9

-150

.9

-149

.9

F-M

g 3

A4)

[%]

48

70

45

97

78

93

92

95

89

89

100

97

97

96

96

91

95

89

100

98

96

FW

HM

3)

[ppm

]

3.6

4.3

4.1

2.7

3.1

3.3

3.3

2.8

3.5

3.6

2.1

2.5

3.1

2.9

2.9

3.1

3.0

3.4

2.5

2.7

2.9

pos.

2)

[ppm

]

-173

.6

-173

.2

-173

.5

-175

.0

-175

.3

-174

.5

-174

.2

-175

.8

-174

.5

-174

.1

-175

.6

-175

.4

-175

.0

-174

.8

-175

.3

-174

.6

-174

.8

-174

.2

-175

.2

-175

.0

-174

.1

x est

1)

0.41

0.51

0.51

0.04

0.11

0.18

0.14

0.24

0.23

0.35

0.00

0.00

0.17

0.07

0.13

0.17

0.29

0.36

0.03

0.03

0.08

y

1.0

1.0

1.0

1.2

1.2

1.2

1.2

1.2

1.2

1.2

1.4

1.4

1.4

1.4

1.4

1.4

1.4

1.4

1.6

1.6

1.6

x

0.6

0.7

0.8

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.1

0.2

0.3

A.2. NMR spectroscopic results

161

Tab

le 6

.4. C

on

tinue

d.

othe

rs

A4)

[%] 2 9 22

1)

x

est =

Add

ition

al A

l-con

tent

of p

hlo

gop

ites

dete

rmin

ed f

rom

29S

i MA

S N

MR

spe

ctra

. Err

or r

ang

e is

± 0

.1

2)

p

os. =

Sig

nal p

ositi

on.

Err

or r

ang

e is

± 0

.5 p

pm.

3)

F

WH

M =

Ful

l wid

th a

t ha

lf m

axim

um E

rror

ran

ge:

± 0

.3 p

pm.

4)

A

= R

elat

ive

sign

al a

rea.

Err

or r

ang

e is

± 3

%.

5)

F

Al/(

FA

l+F

Mg)

= r

elat

ive

F-M

g 2A

l sig

nal

inte

nsity

K3A

lF6*

0.5H

2O A

4)

[%] 9 31

FW

HM

3)

[ppm

]

2.5

2.5

pos.

2)

[ppm

]

-157

.8

-157

.4

FA

l 5)

(FA

l+F

Mg)

0.05

0.04

0.06

0.12

0.17

0.15

0.14

0.03

0.05

0.06

0.07

0.11

0.28

F-M

g 2A

l

A4)

[%]

5 4 6 12

16

5 8 3 5 6 6 10

19

FW

HM

3)

[ppm

]

3.1

3.3

3.2

3.7

3.5

1.7

2.8

2.9

3.1

2.7

2.9

4.0

3.2

pos.

2)

[ppm

]

-149

.9

-150

.3

-149

.7

-149

.3

-149

.1

-147

.8

-150

.0

-150

.5

-149

.5

-149

.9

-149

.7

-149

.3

-147

.4

F-M

g 3

A4)

[%]

95

96

94

88

81

28

50

97

95

94

72

81

50

FW

HM

3)

[ppm

]

3.1

2.8

3.1

3.5

3.6

3.4

3.0

2.9

2.7

3.1

3.2

3.5

3.5

pos.

2)

[ppm

]

-174

.3

-174

.7

-173

.8

-173

.3

-172

.9

-172

.2

-175

.5

-175

.1

-173

.6

-173

.7

-173

.4

-173

.3

-171

.0

x est

1)

0.10

0.35

0.28

0.35

0.63

0.66

0.15

0.17

0.27

0.46

0.71

0.68

y

1.6

1.6

1.6

1.6

1.6

1.6

1.8

1.8

1.8

1.8

1.8

1.8

x

0.4

0.5

0.6

0.8

1.0

1.2

0.2

0.3

0.5

0.6

1.2

1.6

A. Appendix

162

Tab

le 6

.5. P

aram

eter

s ob

tain

ed fr

om 2

7A

l MA

S a

nd M

QM

AS

NM

R s

pect

ra.

[6] A

l

Al 2

O3

A3)

[%]

± 3

CQ

[MH

z]

± 0.

3

FW

HM

CS

2)

[ppm

]

± 0.

5

δ27A

l

[ppm

]

± 0.

5

Phl

ogop

ite

A3)

[%]

± 3

η

± 0.

1

CQ

[MH

z]

± 0.

3

δ27A

l

[ppm

]

± 0.

5

[4] A

l

Phl

ogop

ite

A3)

[%]

± 3

η

± 0.

1

CQ

[MH

z]

± 0.

3

2.4

2.5

2.7

2.7

2.5

2.5

2.4

2.5

2.4

2.4

2.3

2.4

2.4

2.6

2.6

2.5

2.6

2.6

2.5

FW

HM

CS

2)

[ppm

]

± 0.

5

4.4

4.0

4.1

3.4

3.2

3.9

4.3

3.8

4.1

3.8

4.5

4.3

3.9

3.4

3.7

4.5

4.0

4.3

4.3

δ27A

l

[ppm

]

± 0.

5

70.2

69.6

69.5

69.8

70.2

70.1

70.5

70.3

70.6

70.7

70.7

69.7

70.5

70.4

71.3

70.3

70.0

70.1

70.5

Fie

ld

[T]

9.34

x est

1)

± 0.

1

0.24

0.01

0.01

0.00

0.09

0.08

0.29

0.26

0.33

0.33

0.40

0.11

0.37

0.40

0.41

0.10

0.07

0.23

0.16

y

0.0

0.2

0.5

0.5

0.5

0.5

0.5

0.5

0.5

0.5

0.5

0.8

0.8

0.8

0.8

1.0

1.0

1.0

1.0

x

0.8

0.1

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.2

0.4

0.6

0.8

0.1

0.2

0.3

0.4

A.2. NMR spectroscopic results

163

Tab

le 6

.5. C

on

tinue

d.

[6] A

l

Al 2

O3

A3)

[%]

± 3

CQ

[MH

z]

± 0.

3

FW

HM

CS

2)

[ppm

]

± 0.

5

δ27A

l

[ppm

]

± 0.

5

Phl

ogop

ite

A3)

[%]

± 3

η

± 0.

1

CQ

[MH

z]

± 0.

5

δ27A

l

[ppm

]

± 0.

5

[4] A

l

Phl

ogop

ite

A3)

[%]

± 3

η

± 0.

1

CQ

[MH

z]

± 0.

3

2.6

2.6

2.7

2.8

2.6

2.6

2.8

2.6

2.5

2.5

2.8

3.1

2.8

2.7

2.8

2.7

2.4

2.6

2.7

FW

HM

CS

2)

[ppm

]

± 0.

5

3.9

3.8

3.7

4.5

4.1

4.3

4.4

4.2

4.3

4.5

4.0

6.0

4.3

4.3

4.1

4.3

4.6

4.0

3.7

δ27A

l

[ppm

]

± 0.

5

70.6

71.1

71.6

71.7

69.7

70.1

70.6

70.7

70.4

70.8

71.2

68.9

69.6

69.5

70.2

70.0

70.4

70.3

70.8

Fie

ld

[T]

9.3

x est

1)

± 0.

1

0.20

0.41

0.51

0.51

0.04

0.11

0.18

0.14

0.24

0.23

0.35

0.00

0.00

0.17

0.07

0.13

0.17

0.29

0.36

y

1.0

1.0

1.0

1.0

1.2

1.2

1.2

1.2

1.2

1.2

1.2

1.4

1.4

1.4

1.4

1.4

1.4

1.4

1.4

x

0.5

0.6

0.7

0.8

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

A. Appendix

164

Tab

le 6

.5. C

on

tinue

d.

[6] A

l

Al 2

O3

A3)

[%]

± 3

CQ

[MH

z]

± 0.

3

FW

HM

CS

2)

[ppm

]

± 0.

5

δ27A

l

[ppm

]

± 0.

5

Phl

ogop

ite

A3)

[%]

± 3

η

± 0.

1

CQ

[MH

z]

± 0.

3

δ27A

l

[ppm

]

± 0.

5

[4] A

l

Phl

ogop

ite

A3)

[%]

± 3

η

± 0.

1

CQ

[MH

z]

± 0.

3

3.0

2.8

2.6

2.8

2.6

2.6

2.5

2.6

2.6

2.8

2.7

2.6

2.8

2.7

2.7

2.9

2.7

2.9

FW

HM

CS

2)

[ppm

]

± 0.

5

3.9

4.1

4.1

4.3

3.3

3.9

4.4

3.2

3.3

3.7

4.0

4.0

3.1

2.9

3.7

3.3

3.6

3.0

δ27A

l

[ppm

]

± 0.

5

69.9

70.2

70.6

70.8

72.0

71.3

70.7

72.0

72.0

70.2

71.0

70.8

71.8

72.6

72.3

71.0

70.7

72.4

Fie

ld

[T]

9.3

x est

1)

± 0.

1

0.03

0.03

0.08

0.10

0.35

0.28

0.35

0.63

0.66

0.15

0.17

0.27

0.46

0.71

0.68

0.25

0.29

0.67

y

1.6

1.6

1.6

1.6

1.6

1.6

1.6

1.6

1.6

1.8

1.8

1.8

1.8

1.8

1.8

2.0

2.0

2.0

x

0.1

0.2

0.3

0.4

0.5

0.6

0.8

1.0

1.2

0.2

0.3

0.5

0.6

1.2

1.6

0.4

0.5

0.8

A.2. NMR spectroscopic results

165

Tab

le 6

.5. C

on

tinue

d.

[6] A

l

Al 2

O3

A3)

[%]

± 3 25

21

46

46

25

1) x

est =

Add

ition

al A

l-con

tent

of p

hlo

gopi

tes

dete

rmin

ed fr

om

29S

i MA

S N

MR

spe

ctra

.

2) F

WH

M C

S =

Ful

l wid

th a

t hal

f max

imum

of t

he G

auss

ian

chem

ical

shi

ft di

strib

utio

n.

3) A

= R

elat

ive

sign

al in

tens

ity.

CQ

[MH

z]

± 0.

3

5.7

6.4

5.4

5.4

5.7

6.1

6.6

6.2

6.0

FW

HM

CS

2)

[ppm

]

± 0.

5

7.6

7.1

6.9

6.9

7.6 - - - -

δ27A

l

[ppm

]

± 0.

5

17.3

15.4

15.1

15.7

17.3

9.9

8.0

9.4

11.8

Phl

ogop

ite

A3)

[%]

± 3 4 10

5 5 4

η

± 0.

1

0.3

0.3

0.3

0.3

0.3

0.3

0.3

0.3

0.3

CQ

[MH

z]

± 0.

3

4.8

4.7

4.7

4.4

4.8

4.0

3.5

4.0

4.2

δ27A

l

[ppm

]

± 0.

5

12.8

12.1

11.8

11.7

12.8

6.7

7.8

6.7

6.4

[4] A

l

Phl

ogop

ite

A3)

[%]

± 3 71

64

45

44

71

η

± 0.

1

0.6

0.6

0.6

0.6

0.6

0.6

0.6

0.6

0.6

CQ

[MH

z]

± 0.

3

2.5

2.2

2.6

3.7

3.7

3.3

3.1

3.7

2.6

2.6

2.9

1.9

FW

HM

CS

2)

[ppm

]

± 0.

5

2.9

2.6

3.5 - - - - - - - - -

δ27A

l

[ppm

]

± 0.

5

72.6

72.5

72.4

73

74.4

74.5

74.4

73.0

69.8

70.0

69.3

70.9

Fie

ld

[T]

17.6

16.5

x est

1)

± 0.

1

0.83

0.82

0.65

0.25

0.29

0.83

0.82

0.65

0.40

0.51

0.10

0.71

y

2.0

2.0

2.0

2.0

2.0

2.0

2.0

2.0

0.5

1.0

1.6

1.8

x

1.0

1.2

1.6

0.4

0.8

1.0

1.2

1.6

0.8

0.8

0.4

1.2

Al 2

O3

A. Appendix

166

Table 6.6. {1H} → 29Si contact time dependent magnetisation curves. A = overall intensity of Si-nAl signals. I = point of highest intensity of the Si-Si2Al or Si-SiAl2 signal.

x = 0.4 y = 1.0

x = 0.5 y = 1.0

x = 0.6 y = 1.0

x = 0.7 y = 1.0

x = 0.8 y = 1.0

800 °C 800 °C 800 °C 800 °C 800 °C

contact time [ms]

A [1010 a.u.]

I [108 a.u.]

A [1010 a.u.]

I [108 a.u.]

A [1010 a.u.]

I [108 a.u.]

A [1010 a.u.]

I [108 a.u.]

A [1010 a.u.]

I [108 a.u.]

0.1 0.0 0.0 0.1 0.0 0.0 0.0 0.2 0.0 0.1 0.2

0.2 0.2 0.2 0.2 0.2 0.3 0.2 0.2 0.3 0.3 0.3

0.3 0.1 0.2 0.5 0.4 0.3 0.4 0.3 0.3 0.4 0.4

0.4 0.3 0.3 0.4 0.5 0.4 0.4 0.4 0.4 0.4 0.4

0.5 0.3 0.3 0.5 0.6 0.6 0.5 0.5 0.5 0.5 0.5

1.0 0.6 0.6 0.9 0.9 0.9 0.8 0.9 0.7 0.9 0.9

1.5 0.8 0.8 1.2 1.2 1.3 1.1 1.1 0.9 1.2 1.0

2.0 0.9 0.9 1.4 1.4 1.5 1.2 1.4 1.1 1.4 1.2

2.5 1.0 1.0 1.6 1.7 1.5 1.4 1.6 1.3 1.6 1.4

3.0 1.1 1.2 1.7 1.7 1.6 1.4 1.8 1.3 1.7 1.5

3.5 1.2 1.3 1.8 1.9 1.6 1.5 1.7 1.4 1.9 1.7

4.0 1.2 1.3 2.0 2.0 1.7 1.5 2.0 1.6 1.9 1.7

5.0 1.5 1.5 2.0 2.2 2.0 1.6 2.2 1.8 2.1 1.8

6.0 1.4 1.6 2.2 2.4 2.0 1.8 2.3 1.9 2.3 2.0

7.0 1.5 1.7 2.3 2.5 2.1 1.9 2.6 2.0 2.5 2.0

8.0 1.7 1.8 2.5 2.6 2.3 2.0 2.7 2.3 2.6 2.2

9.0 1.7 1.9 2.5 2.8 2.2 2.1 2.8 2.4 2.8 2.3

10 1.8 2.1 2.7 2.9 2.4 2.1 2.9 2.4 2.8 2.5

15 2.0 2.3 3.1 3.3 2.7 2.5 3.4 2.9 3.4 2.9

20 2.1 2.3 3.3 3.7 3.1 2.8 3.6 3.1 3.6 3.1

25 2.1 2.4 3.5 3.8 3.1 2.9 3.7 3.2 3.9 3.4

30 2.3 2.6 3.5 4.0 3.2 3.1 3.8 3.3 4.1 3.7

40 2.4 2.7 3.8 4.1 3.3 3.2 4.0 3.5 4.3 3.8

50 2.5 2.7 3.9 4.2 3.4 3.3 4.1 3.5 4.5 4.0

60 2.5 2.8 3.9 4.3 3.7 3.4 4.0 3.7 4.6 4.1

80 2.5 2.8 3.9 4.4 3.7 3.5 4.0 3.6 4.6 4.1

100 2.5 2.8 3.8 4.4 3.8 3.5 4.2 3.7 4.5 4.1

120 2.5 2.7 4.0 4.4 3.7 3.5 4.2 3.7 4.4 4.1

A.2. NMR spectroscopic results

167

Table 6.6. Continued.

x = 0.4 y = 1.5

x = 0.6 y = 1.5

x = 0.8 y = 1.8

x = 0.5 y = 1.6

x = 0.7 y = 1.8

800 °C 800 °C 800 °C 600 °C 600 °C

contact time [ms]

A [1010 a.u.]

I [108 a.u.]

A [1010 a.u.]

I [108 a.u.]

A [1010 a.u.]

I [108 a.u.]

A [1010 a.u.]

I [108 a.u.]

A [1010 a.u.]

I [108 a.u.]

0.1 0.2 0.2 0.1 0.2 0.2 0.2 0.1 0.0 0.2 0.0

0.2 0.4 0.4 0.2 0.3 0.5 0.5 0.3 0.3 0.3 0.2

0.3 0.5 0.5 0.5 0.5 0.7 0.6 0.5 0.4 0.4 0.4

0.4 0.8 0.7 0.4 0.6 0.8 0.7 0.6 0.5 0.6 0.5

0.5 0.8 0.8 0.5 0.7 0.9 0.8 0.7 0.6 0.7 0.6

1.0 1.4 1.1 0.9 0.9 1.1 1.1 1.0 0.9 1.2 1.0

1.5 1.5 1.3 1.2 1.1 1.4 1.5 1.2 1.1 1.6 1.2

2.0 1.7 1.4 1.4 1.2 1.7 1.7 1.3 1.2 1.8 1.4

2.5 1.9 1.6 1.6 1.5 1.6 1.7 1.6 1.4 1.9 1.5

3.0 2.0 1.7 1.7 1.5 1.8 1.8 1.7 1.5 2.0 1.6

3.5 2.2 1.9 1.8 1.6 1.8 1.9 1.7 1.6 2.2 1.7

4.0 2.1 1.9 2.0 1.7 1.9 2.0 1.8 1.6 2.2 1.8

5.0 2.1 1.9 2.0 1.8 2.0 2.1 1.9 1.7 2.2 1.8

6.0 2.3 2.0 2.2 1.8 2.2 2.3 2.0 1.8 2.3 1.9

7.0 2.6 2.1 2.3 1.9 2.2 2.4 2.1 1.9 2.4 1.9

8.0 2.7 2.3 2.5 2.0 2.4 2.5 2.2 2.0 2.5 2.0

9.0 2.8 2.4 2.5 2.1 2.5 2.5 2.3 2.1 2.5 2.1

10 2.6 2.4 2.7 2.2 2.3 2.6 2.4 2.2 2.6 2.1

15 3.1 2.7 3.1 2.4 2.7 2.9 2.6 2.4 2.7 2.2

20 3.2 2.9 3.5 2.6 3.0 3.1 2.8 2.6 2.8 2.3

25 3.5 3.1 3.5 2.7 3.1 3.3 2.9 2.7 3.0 2.5

30 3.6 3.3 3.5 2.6 3.3 3.4 3.1 2.9 3.1 2.5

40 3.9 3.5 3.8 2.8 3.4 3.5 3.1 3.0 3.2 2.7

50 4.1 3.7 3.9 2.9 3.5 3.7 3.2 3.0 3.3 2.8

60 4.0 3.6 3.9 3.0 3.5 3.7 3.2 3.0 3.2 2.7

80 4.2 3.8 3.9 3.0 3.6 3.7 3.3 3.1 3.2 2.7

100 4.3 3.9 3.8 3.0 3.8 3.9 3.3 3.1 3.2 2.8

120 4.4 4.0 4.0 3.0 3.6 3.8 3.2 3.1 3.2 2.8

A. Appendix

168

Table 6.7. {19F} → 29Si contact time dependent magnetisation curves. A = overall intensity of Si-nAl signals. I = point of maximum intensity of the Si-Si2Al or Si-SiAl2 signal.

x = 0.1/y = 0.2 x = 0.0/y = 0.5 x = 0.3/y = 0.5

contact time [ms]

A [109 a.u.]

I [108 a.u.]

A [1010 a.u.]

I [108 a.u.]

A [1010 a.u.]

I [108 a.u.]

0.1 0.2

0.2 1.2 0.1 0.4

0.3 0.1 0.3 0.7 0.2

0.4 0.2 0.0 0.7 0.2

0.5 0.3 0.8 1.1 0.3

1.0 0.8 0.5 3.3 0.6

1.5 0.4 0.5 0.2 6.2 0.7

2.0 1.4 0.2 1.9 0.3 8.4 1.0

2.5 1.8 0.3 2.8 0.4 9.2 1.3

3.0 2.4 0.4 2.3 0.4 10.5 1.4

3.5 3.9 0.4 2.6 0.5 12.2 1.5

4.0 4.2 0.5 3.7 0.6 13.4 1.6

5.0 5.6 0.8 5.6 0.9 15.3 2.0

6.0 7.5 0.9 7.5 1.2 16.8 2.0

7.0 11.2 1.3 8.0 1.3 17.2 2.2

8.0 12.3 1.6 9.8 1.6 18.2 2.3

9.0 13.8 1.7 11.2 1.8 19.8 2.4

10 17.4 2.1 11.8 1.9 21.8 2.6

15 23.7 3.1 15.9 2.6 24.1 2.7

20 29.2 3.7 20.8 3.4 23.4 2.9

25 34.7 4.5 22.2 3.8 23.8 3.0

30 37.1 4.7 23.3 4.0 25.8 3.0

40 35.6 4.6 25.2 4.1 26.2 3.2

50 33.2 4.1 22.9 3.7 26.3 3.3

60 31.6 4.0 20.3 3.4 25.0 3.1

80 26.9 3.5 15.3 2.7 26.9 3.2

100 20.2 2.5 11.1 1.9 23.9 3.0

120 9.8 1.4 7.6 1.4 25.1 3.0

A.2. NMR spectroscopic results

169

Table 6.7. Continued.

x = 0.6/y = 0.5 x = 0.8/y = 0.8 x = 0.7/y = 1.0

contact time [ms]

A [1010 a.u.]

I [108 a.u.]

A [1010 a.u.]

I [108 a.u.]

A [1010 a.u.]

I [108 a.u.]

0.1 0.1 1.0

0.2 0.4 0.5 0.2

0.3 1.4 0.2 6.8 1.4 0.2

0.4 2.0 0.2 4.8 0.7 2.4

0.5 3.8 0.2 6.2 2.4

1.0 5.9 0.6 7.7 0.7 4.2

1.5 7.1 0.7 7.5 6.2 0.5

2.0 10.5 0.9 11.0 1.3 7.8 0.6

2.5 10.6 1.1 13.0 8.5 0.6

3.0 14.0 1.2 18.2 1.4 11.7 0.8

3.5 14.7 1.3 16.8 1.5 11.5 1.0

4.0 18.4 1.4 20.0 1.7 13.5 1.0

5.0 19.1 1.8 22.8 2.0 12.9 1.1

6.0 21.5 2.1 24.2 2.0 14.6 1.3

7.0 26.7 2.4 24.4 2.4 16.4 1.4

8.0 26.4 2.5 27.7 2.4 15.2 1.5

9.0 28.9 2.6 33.2 2.6 20.4 1.5

10 29.6 2.7 32.1 2.7 18.6 1.5

15 35.4 3.2 33.8 3.2 22.9 2.0

20 42.3 3.6 36.9 3.5 24.8 2.1

25 40.2 3.6 47.9 3.9 27.3 2.3

30 42.0 3.8 37.8 3.8 26.7 2.4

40 39.2 3.7 42.0 4.0 30.5 2.7

50 40.0 3.6 46.0 4.3 29.5 2.7

60 41.3 3.4 40.5 4.2 29.8 2.6

80 40.3 3.4 50.2 4.7 32.9 2.8

100 33.5 3.0 45.2 4.1 31.6 2.5

120 35.6 2.9 40.7 4.2 30.6 2.5

B. References

171

B. References

ABRAGAM, A. (2007) Principles of nuclear magnetism. Oxford University Press, Oxford.

ALIETTI, E., Brigatti, M.F., Poppi, L. (1995) The crystal-structure and chemistry of high-aluminum phlogopite. Mineralogical Magazine, 59, 149-157.

AMOUREUX, J.-P., Fernandez, C. Triple, quintuple and higher order multiple quantum MAS NMR of quadrupolar nuclei. Solid State Nuclear Magnetic Resonance, 10, 211-223.

ARTACHO, E., Sánchez-Portal, D., Ordejón, P., García, A., Soler, J.M. (1999) Linear-scaling ab-initio calculations for large and complex systems. Physica Status Solidi (b), 215, 809-817.

BAILEY, S.W. (1975) Cation ordering and pseudo-symmetry in layer silicates. American Mineralogist, 60, 175-187.

BAILEY, S.W. (1984a) Review of cation ordering in micas. Clays and clay minerals, 32, 81-92.

BAILEY, S.W. (1984b) Micas. Reviews in Mineralogy and Geochemistry, 13, Mineralogical Society of America, Washington, D.C.

BAKER, D.R., Bossànyi, H. (1994) The combined effect of F and H2O on interdiffusion between peralkaline dacitic and rhyolitic melts. Contributions to Mineralogy and Petrology, 117, 203-214.

BAKER, D.R., Vaillancourt, J. (1995) The low viscosities of F + H2O-bearing granitic melts and implications for melt extraction and transport. Earth and Planetary Science Letters, 132, 199-211.

BIGI, S., Brigatti, M.F., Mazzucchelli, M., Rivalenti, G. (1993) Crystal chemical variations in Ba-rich biotites from gabbroic rocks of lower crust (Ivrea zone, NW Italy). Contributions to Mineralogy and Petrology, 113, 87-99.

BIGI, S., Brigatti, M.F. (1994) Crystal-chemistry and microstructures of plutonic biotite. American Mineralogist, 79, 63-72.

BOSENICK, A., Dove, M.T., Myers, E.R., Palin, E.J., Sainz-Diaz, C.I., Guiton, B.S., Warren, M.C., Craig, M.S., Redfern, S.A.T. (2001) Computational methods for the study of energies of cation distributions: applications to cation-ordering phase transitions and solid solutions. Mineralogical Magazine, 65, 193-219.

BOUKILI, B., Robert, J.-L., Beny, J.-M., Holtz, F. (2001) Structural effects of OH – F substitution in trioctahedral micas of the system: K2O-FeO-Fe2O3-Al2O3-SiO2-H2O-HF. Schweizerische Mineralogische und Petrographische Mitteilungen, 81, 55-67.

BRIGATTI, M.F., Frigieri, P., Ghezzo, C., Poppi, L. (2000) Crystal chemistry of Al-rich biotites coexisting with muscovites in peraluminous granites. American Mineralogist, 85, 436-448.

B. References

172

BRIGATTI, M.F., Guggenheim, S. (2002) Mica crystal chemistry and the influence of pressure, temperature, and solid solution on atomistic models. In: Mottana, A., Sassi, F.P., Thompson Jr., J.B., Guggenheim, S. (Eds.), Micas: Crystal chemical and metamorphic petrology, Reviews in Mineralogy and Geochemistry, 46, 1-97. Mineralogical Society of America, Washington, D.C.

CARROLL, M.R., Webster, J.D. (1994) Solubilities of sulphur, noble-gases, nitrogen, chlorine, and fluorine in magmas. In: Carroll, M.R., Holloway, J.R. (Eds.), Volatiles in Magmas, Reviews in Mineralogy and Geochemistry, 30, 231-279. Mineralogical Society of America, Washington, D.C.

CHRISTIANSEN, E.H., Burt, D.M., Sheridan, M.F., Wilson, R.T. (1983) The petrogenesis of topaz rhyolites from the Western United States. Contributions to Mineralogy and Petrology, 83, 16-30.

CIRCONE, S., Navrotsky, A., Kirkpatrick, R.J., Graham, C.M. (1991) Substitution of [6,4]Al in phlogopite: Mica characterization, unit-cell variation, 27Al and 29Si MAS-NMR spectroscopy, and Al-Si distribution in the tetrahedral sheet. American Mineralogist, 76, 1485-1501.

CIRCONE, S., Navrotsky, A. (1992) Substitution of [6,4]Al in phlogopite: High-temperature solution calorimetry, heat capacities, and thermodynamic properties of the phlogopite-eastonite join. American Mineralogist, 77, 1191-1205.

CRUCIANI, G., Zanazzi, P.F. (1994) Cation partitioning and substitution mechanisms in 1M phlogopite – a crystal-chemical study. American Mineralogist, 79, 289-301.

DINGWELL, D.B., Scarfe, C.M., Cronin, D.J. (1985) The effect of fluorine on viscosities in the system Na2O-Al2O3-SiO2: implications for phonolites, trachytes and rhyolites. American Mineralogist, 70, 80-87.

DINGWELL, D.B., Knoche, R., Webb, S.L. (1993) The effect of F on the density of haplogranite melt. American Mineralogist, 78, 325-330.

DUER, M.J. (2002) The basics of solid-state NMR. In: Duer, M.J. (Ed.), Solid-state NMR spectroscopy – principles and applications, Blackwell Science, Oxford, 3-72.

ĎUROVIČ, S. (1994) Classification of phillosilicates according to the symmetry of their octahedral sheets. Ceramics-Silikáty, 38, 81-84.

DUŠEK, M., Petříček, V., Wunschel, M., Dinnebier, R.E., van Smaalen, S. (2001) Refinement of modulated structures against X-ray powder diffraction data with JANA2000. Journal of Applied Crystallography, 34, 398-404.

ENDO, Y. (1968) Growth spirals and polytypism of fluor-phlogopite. Journal of the Mineralogical Society of Japan, 8, Spec. Issue 2, 39-41.

B. References

173

ERNST, R.R., Bodenhausen, G., Wokaun, A. (1994) Principles of nuclear magnetic resonance in one and two dimensions. Oxford University Press, Oxford.

FECHTELKORD, M., Behrens, H., Holtz, F., Fyfe, C.A., Groat, L.A., Raudsepp, M. (2003a) Influence of F content on the composition of Al-rich synthetic phlogopite: Part I. New information on structure and phase-formation from 29Si, 1H, and 19F MAS NMR spectroscopies. American Mineralogist, 88, 47-53.

FECHTELKORD, M., Behrens, H., Holtz, F., Bretherton, J.L., Fyfe, C.A., Groat, L.A. Raudsepp, M. (2003b) Influence of F content on the composition of Al-rich synthetic phlogopite: Part II. Probing the structural arrangement of aluminum in tetrahedral and octahedral layers by 27Al MQMAS and 1H/19F-27Al HETCOR and REDOR experiments. American Mineralogist, 88, 1046-1057.

FERRARIS, G., Gula, A., Ivaldi, G., Nespolo, M., Sokolova, E., Uvarova, Y., Khomyakov, A.P. (2001) First structure determination of an MDO-2O mica polytype associated with a 1M polytype. European Journal of Mineralogy, 13, 1013-1023.

FERRARIS, G., Ivaldi, G. (2002) Structural features of micas. In: Mottana, A., Sassi, F.P., Thompson Jr., J.B., Guggenheim, S. (Eds.), Micas: Crystal chemical and metamorphic petrology, Reviews in Mineralogy and Geochemistry, 46, 117-153. Mineralogical Society of America, Washington, D.C.

FOSTER, M.D. (1960) Interpretation of the composition of lithium micas. Geological Survey Professional Paper, 354-E, 115-147.

FREGOLA, R.A., Capitani, G., Scandale, E., Ottolini, L. (2009) Chemical control of 3T stacking order in a Li-poor biotite mica. American Mineralogist, 94, 334-344.

FRYDMAN, L., Harwood, J.S. (1995) Isotropic spectra of half-integer quadrupolar spins from bidimensional magic-angle spinning NMR. Journal of the American Chemical Society, 117, 5367-5368.

GALE, J.D. (1997) GULP: A computer program for the symmetry-adapted simulation of solids. Journal of the Chemical Society: Faraday Transactions, 93, 629-637.

GIORDANO, D., Romano, C., Dingwell, D.B., Poe, B., Behrens, H. (2004) The combined effects of water and fluorine on the viscosity of silicic magmas. Geochimica et Cosmochimica Acta, 68, 5159-5168.

GREEN, T.H. (1981) Synthetic high-pressure micas compositionally intermediate between the dioctahedral and trioctahedral mica series. Contributions to Mineralogy and Petrology, 78, 452-458.

HAMILTON, D.L., Henderson, C.M.B. (1968) The preparation of silicate compositions by a gelling method. Mineralogical Magazine, 36, 832-838.

HARRIS, R.K., Jackson, P. (1991) High-resolution F-19 magnetic-resonance of solids. Chemical Reviews, 91, 1427-1440.

B. References

174

HARTMANN, S.R., Hahn, E.L. (1962) Nuclear double resonance in the rotating frame. Physical Review, 128, 2042-2053.

HAYASHI, S., Akiba, E. (1994) Interatomic distances in layered silicates and their intercalation compounds as studied by cross polarization NMR. Chemical Physics Letters, 226, 495-500.

HENDRICKS, B.S., Jefferson, M.E. (1939) Polymorphism of the micas. American Mineralogist, 24, 729-771.

HERRERO, C.P., Sanz, J., Serratosa, J.M. (1985a) Si, Al distribution in micas: Analysis by high-resolution 29Si NMR spectroscopy. Journal of Physics C: Solid State Physics, 19, 4169-4181.

HERRERO, C.P., Sanz, J., Serratosa, J.M. (1985b) Tetrahedral cation ordering in layer silicates by 29Si NMR spectroscopy. Solid State Communications, 53, 151-154.

HERRERO, C.P., Gregorkiewitz, M., Sanz, J., Serratosa, J.M. (1987) 29Si MAS-NMR spectroscopy of mica-type silicates: Observed and predicted distribution of tetrahedral Al-Si. Physics and Chemistry of Minerals, 15, 84-90.

HEWITT, D.A., Wones, D.R. (1975) Physical properties of some synthetic Fe-Mg-Al trioctahedral biotites. American Mineralogist, 60, 854-862.

HUVE, L., Delmotte, L., Martin, P., Le Dred, R., Baron, J., Saehr, D. (1992a) 19F MAS-NMR study of structural fluorine in some natural and synthetic 2:1 layer silicates. Clays and Clay Minerals, 40, 186-191.

HUVE, L., Saehr, D., Delmotte, L., Baron, J., Le Dred, R. (1992b) Spectrométrie de Résonance Magnétique Nucléaire du fluor-19 de phyllosilicates et phyllogermanates fluorés. Comptes rendus de l’Académie des Sciences Paris, 315, Série II, 545-549.

KNOCHE, R., Dingwell, D.B., Webb, S.L. (1995) Melt densities for leucogranites and granitic pegmatites: partial molar volumes for SiO2, Al2O3, Na2O, K2O, Li2O, Rb2O, Cs2O, MgO, CaO, SrO, BaO, B2O3, P2O5, F2O-1, TiO2, Nb2O5, Ta2O5, and WO3. Geochimica et Cosmochimica Acta, 59, 4645-4652.

LANGER, K., Chatterjee, N.D., Abraham, K. (1981) Infrared studies of some synthetic and natural 2M1 dioctahedral micas. Neues Jahrbuch für Mineralogie - Abhandlungen, 142, 91-110.

LEE, S.K., Stebbins, J.F. (2003a) O atom sites in natural kaolinite and muscovite: 17O MAS and 3QMAS NMR study. American Mineralogist, 88, 493-500.

LEE, S.K., Stebbins, J.F. (2003b) 17O and 27Al MAS and 3QMAS NMR study of synthetic and natural layer silicates. Chemistry of Materials, 15, 2605-2613.

LIPSICAS, M., Raythatha, R.H., Pinnavaia, T.J., Johnson, I.D., Giese, R.F., jr., Costanzo, P.M., Robert, J.L. (1984) Silicon and aluminium site distributions in 2:1 layered silicate clays. Nature, 309, 604-607.

B. References

175

LOEWENSTEIN, W. (1954) The distribution of aluminum in the tetrahedra of silicates and aluminates. American Mineralogist, 39, 92-96.

LONDON, D. (1987) Internal differentiation of rare-element pegmatites – effects of boron, phosphorus, and fluorine. Geochimica et Cosmochimica Acta, 51, 403-420.

LUTH, W.C., Tuttle, O.F. (1963) Externally heated cold-seal pressure vessels for use to 10,000 bars and 750 degrees C. American Mineralogist, 48, 1401-1403.

MANNING, D.A.C. (1981) The effect of fluorine on liquidus phase relationships in the system Qz-Ab-Or with excess water at 1 kb. Contributions to Mineralogy and Petrology, 76, 206-215.

MARICQ, M.M., Waugh, J.S. (1979) NMR in rotating solids. Journal of Chemical Physics, 70, 3300-3316.

MASSIOT, D., Touzo, B., Trumeau, D., Coutures, J.P., Virlet, J., Florian, P., Grandinetti, P.J. (1996) Two-dimensional magic-angle spinning isotropic reconstruction sequences for quadrupolar nuclei. Solid State Nuclear Magnetic Resonance, 6, 73-83.

MASSIOT, D., Fayon, F., Capron, M., King, I., Le Calvé, S., Alonso, B., Durand, J.-O., Bujoli, B., Gan, Z., Hoatson, G. (2002) Modelling one- and two-dimensional solid-state NMR spectra. Magnetic Resonance in Chemistry, 40, 70-76.

MCCAULEY, J.W., Newnham, R.E., Gibbs, G.V. (1973) Crystal-structure analysis of synthetic fluorophlogopite. American Mineralogist, 58, 249-254.

MEDEK, A., Harwood, J.S., Frydman, L. (1995) Multiple-quantum magic-angle spinning NMR: A new method for the study of quadrupolar nuclei in solids. Journal of the American Chemical Society, 117, 12779-12787.

MEDEK, A., Marinelli, L., Frydman, L. (1998) Multiple-quantum magic-angle spinning NMR of half-integer quadrupolar nuclei. In: Fitzgerald, J.J. (Ed.), Solid-state NMR spectroscopy of inorganic materials, 136-155. American Chemical Society Symposium Series, American Chemical Society, Washington, D.C.

MERCIER, P.H.J., Rancourt, D.G., Robert, J.-L., Berman, R.G., Redhammer, G.J. (2005) Fundamental difference between synthetic powder and natural or synthetic single-crystal 1M micas: Geometric homo-octahedral vs. geometric meso-octahedral sheets. American Mineralogist, 90, 399-410.

MILLOT, Y., Man, P.P. (2002) Procedures for labeling the high-resolution axis of two-dimensional MQ-MAS NMR spectra of half-integer quadrupole spins. Solid State Nuclear Magnetic Resonance, 21, 21-43.

MOTTANA, A., Sassi, F.P., Thompson, J.B., jr., Guggenheim, S. (2002) Micas: Crystal chemistry and metamorphic petrology. Reviews in Mineralogy and Geochemistry, 46, Mineralogical Society of America, Washington, D.C.

MÜLLER, R.F. (1972) Stability of biotite: A discussion. American Mineralogist, 57, 300-316.

B. References

176

MÜLLER, D., Hoebbel, D., Gessner, W. (1981) 27Al NMR studies of aluminosilicate solutions. Influences of the second coordination sphere on the shielding of aluminium. Chemical Physics Letters, 84, 25-27.

MUELLER, K.T., Sun, B.Q., Chingas, G.C., Zwanziger, J.W., Terao, T., Pines, A. (1990) Dynamic-angle spinning of quadrupolar nuclei. Journal of Magnetic Resonance, 86, 470-487.

NESPOLO, M., Ďurovič, S. (2002) Crystallographic basis of polytypism and twinning in micas. In: Mottana, A., Sassi, F.P., Thompson Jr., J.B., Guggenheim, S. (Eds.), Micas: Crystal chemical and metamorphic petrology, Reviews in Mineralogy and Geochemistry, 46, 155-279. Mineralogical Society of America, Washington, D.C.

ORDEJÓN, P., Artacho, E., Soler, J.M. (1996) Self-consistent order-N density-functional calculations for very large systems. Physical Reviews B, 15, 10441-10444.

PALIN, E.J., Dove, M.T., Redfern, S.A.T., Bosenick, A., Sainz-Díaz, C.I., Warren, M.C. (2001) Computational study of tetrahedral Al-Si ordering of muscovite. Physics and Chemistry of Minerals, 28, 534-544.

PALIN, E.J., Dove, M.T., Redfern, S.A.T., Sainz-Díaz, C.I., Lee, W.T. (2003) Computational study of tetrahedral Al-Si and octahedral Al-Mg ordering in phengite. Physics and Chemistry of Minerals, 30, 293-304.

PALIN, E.J., Dove, M.T. (2004) Investigation of Al/Si ordering in tetrahedral phyllosilicate sheets by Monte Carlo simulation. American Mineralogist, 89, 176-184.

PALIN, E.J., Harrison, R.J. (2007) A computational investigation of cation ordering phenomena in the binary spinel system MgAl2O4-FeAl2O4. Mineralogical Magazine, 71, 611-624.

PAPIN, A., Sergent, J., Robert, J.-L. (1997) Intersite OH-F distribution in an Al-rich synthetic phlogopite. European Journal of Mineralogy, 9, 501-509.

PAULING, L. (1960) The nature of the chemical bond. Cornell University Press, Ithaca, New York.

PICHAVANT, M., Manning, D.A.C. (1984) Petrogenesis of tourmaline granites and topaz granites – the contribution of experimental data. Physics of the Earth and Planetary Interiors, 35, 31-50.

PINES, A., Gibby, M.G., Waugh, J.S. (1972) Proton-enhanced nuclear induction spectroscopy. A method for high resolution NMR of dilute spins in solids. Journal of Chemical Physics, 56, 1776-1777.

PINES, A., Gibby, M.G., Waugh, J.S. (1973) Proton-enhanced NMR of dilute spins in solids. Journal of Chemical Physics, 59, 569-590.

PINI, S., Affronte, M., Brigatti, M.F. (2008) Magnetic behaviour of trioctahedral mica-2M1 occurring in a magnetic anomaly zone. Mineralogical Magazine, 72, 1035-1042.

B. References

177

RAMSDELL, L.S. (1947) Studies on silicon carbide. American Mineralogist, 32, 64-82.

RAO, N.V.C., Srivastava, R.K. (2009) Petrology and geochemistry of diamondiferous Mesoproterozoic kimberlites from Wajrakarur kimberlite field, Eastern Dhwarwar craton, southern India: genesis and constraints on mantle source regions. Contributions to Mineralogy and Petrology, 157, 245-265.

RICKWOOD, P.C. (1981) The largest crystals. American Mineralogist, 66, 885-907.

RIEDER, M., Cavazzini, G., D’Yakonov, Y.S., Frank-Kamenetskii, V.A., Gottardi, G., Guggenheim, S., Koval, P.V., Müller, G., Neiva, A.M.R., Radoslovich, E.W., Robert, J.-L., Sassi, F.P., Takeda, H., Weiss, Z., Wones, D.R. (1998) Nomenclature of the micas. The Canadian Mineralogist, 36, 905-912.

ROBERT, J.-L. (1976) Phlogopite solid solutions in the system K2O-MgO-Al2O3-SiO2-H2O. Chemical Geology, 17, 195-212.

ROBERT, J.-L., Kodama, H. (1988) Generalization of the correlations between hydroxyl-stretching wavenumbers and composition of micas in the system K2O-MgO-Al2O3-SiO2-H2O: A single model for trioctahedral and dioctahedral micas. American Journal of Science, 288-A, 196-212.

ROBERT, J.-L., Beny, J.-M., Della Ventura, G., Hardy, M. (1993) Fluorine in micas: crystal-chemical control of the OH-F distribution between trioctahedral and dioctahedral sites. European Journal of Mineralogy, 5, 7-18.

RUTHERFORD, M.J. (1973) The phase relations of aluminous iron biotites in the system KAlSi3O8-KAlSiO4-Al2O3-Fe-O-H. Journal of Petrology, 14, 159-180.

SAINZ-DÍAZ, C.I., Palin, E.J., Hernández-Laguna, A., Dove, M.T. (2003a) Octahedral cation ordering of illite and smectite. Theoretical exchange potential determination and Monte Carlo simulations. Physics and Chemistry of Minerals, 30, 382-392.

SAINZ-DÍAZ, C.I., Palin, E.J., Dove, M.T., Hernández-Laguna, A. (2003b) Monte Carlo simulations of ordering of Al, Fe, and Mg cations in the octahedral sheet of smectites and illites. American Mineralogist, 88, 1033-1045.

SAMOSON, A., Lippmaa, E., Pines, A. (1988) High resolution solid-state N.M.R. Averaging of second-order effects by means of a double-rotor. Molecular Physics, 65, 1013-1018.

SANZ, J., Serratosa, J.M. (1984) 29Si and 27Al high-resolution MAS-NMR spectra of phyllosilicates. Journal of the American Chemical Society, 106, 4790-4793.

SCHINGARO, E., Scordari, F., Ventruti, G. (2001) Trioctahedral micas-1M from Mt. Vulture (Italy): Structural disorder and crystal chemistry. European Journal of Mineralogy, 13, 1057-1069.

SCHREYER, W., Abraham, K., Kulke, H. (1980) Natural sodium phlogopite coexisting with potassium phlogopite and sodium aluminium talc in a metamorphic evaporite sequence from Derrag, Tell Atlas, Algeria. Contributions to Mineralogy and Petrology, 74, 223-233.

B. References

178

SHANNON, R.D. (1976) Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallographica A, 32, 751-767.

SHELL, H.R., Ivey, K.H. (1969) Fluorine micas. U.S. Department of the Interior, Bureau of Mines, 647, 291 p.

SLICHTER, C.P. (1996) Principles of magnetic resonance. Springer, Berlin.

SOLER, J.M., Artacho, E., Gale, J.D., García, A., Junquera, J., Ordejón, P., Sánchez-Portal, D. (2002) The SIESTA method for ab initio order-N materials simulation. Journal of Physics: Condensed Matter, 14, 2745-2779.

STEBBINS, J.F., Zhao, P., Lee, S.K., Cheng, X. (1999) Reactive Al-O-Al stes in a natural zeolite: Triple-quantum oxygen-17 nuclear magnetic reasonance. American Mineralogist, 84, 1680-1684.

SUNAGAWA, I., Endo, J., Daimon, N., Tate, I. (1968) Nucleation, growth and polytypism of fluor-phlogopite from the vapour phase. Journal of Crystal Growth, 3-4, 751.

TAKEDA, H., Morosin, B. (1975) Comparison of observed and predicted structural parameters of mica at high-temperature. Acta Crystallographica B, 31, 2444-2452.

TATEYAMA, H., Shimoda, S., Sudo, T. (1974) The crystal structure of synthetic Mg(IV)mica. Zeitschrift für Kristallographie, Kristallgeometrie, Kristallphysik, Kristallchemie, 139, 196-206.

TISCHENDORF, G., Förster, H.-J., Gottesmann, B., Rieder, M. (2007) True and brittle micas: composition and solid-solution series. Mineralogical Magazine, 71, 285-320.

TORAYA, H. (1981) Distortions of octahedra and octahedral sheets in 1M micas and the relation to their stability. Zeitschrift für Kristallographie, 157, 173-190.

TUTTLE, O.F. (1949) Two pressure vessels for silicate-water studies. Bulletin of the Geological Society of America, 60, 1727-1729.

VAN ECK, E.R.H., Smith, M.E., Kohn, S.C. (1999) Observation of hydroxyl groups by 17O solid-state multiple quantum MAS NMR in sol-gel-produced silica. Solid State Nuclear Magnetic Resonance, 15, 181-188.

VAN VALKENBURG, A., Pike, R.G. (1952) Synthesis of mica. Journal of Research of the National Bureau of Standards, 48, 360-369.

VAN VLECK, J.H. (1948) The dipolar broadening of magnetic resonance lines in crystals. Physical Review, 74, 1168-1183.

VIRGO, D, Popp, R.K. (2000) Hydrogen deficiency in mantle-derived phlogopites. American Mineralogist, 85, 753-759.

B. References

179

WALTHER, K.L., Wokaun, A., Baiker, A. (1990) Characterization of porous silica gels prepared via the sol-gel process by 29Si CP/MAS solid-state NMR spectroscopy. Molecular Physics, 71, 769-780.

WARREN, M.C., Dove, M.T., Myers, E.R., Bosenick, A., Palin, E.J., Sainz-Diaz, C.I., Guiton, B.S., Redfern, S.A.T. (2001) Monte Carlo methods for the study of cation ordering in minerals. Mineralogical Magazine, 65, 221-248.

WEBSTER, J.D., Holloway, J.R., Hervig, R.L. (1987) Phase equilibria of a Be, U and F-enriched vitrophyre from Spor Mountain, Utah. Geochimica et Cosmochimica Acta, 51, 389-402.

WEIDNER, J.R., Martin, R.F. (1987) Phase equilibria of a F-rich leucogranite from the St. Austell pluton, Cornwall. Geochimica et Cosmochimica Acta, 51, 1591-1597.

WEISS, C.A., Jr., Altaner, S.P., Kirkpatrick, R.J. (1987) High-resolution 29Si NMR spectroscopy of 2:1 layer silicates: Correlations among chemical shift, structural distortions, and chemical variations. American Mineralogist, 72, 935-942.

WEISS, Z., Rieder, M., Chmielovà, M. (1992) Deformation of coordination polyhedra and their sheets in phillosilicates. European Journal of Mineralogy, 4, 665-682.

WHALEN, J.B., Currie, K.L., Chappell, B.W. (1987) A-type granites: geochemical characteristics, discrimination, and petrogenesis. Contributions to Mineralogy and Petrology, 95, 407-419.

WOESSNER, D.E. (1989) Characterization of clay minerals by 27Al nuclear magnetic resonance spectroscopy. American Mineralogist, 74, 203-215.

WONES, D.R., Eugster, H.P. (1965) Stability of biotite: Experiment, theory, and application. American Mineralogist, 50, 1228-1272.

WONES, D.R. (1967) A low pressure investigation of the stability of phlogopite. Geochimica et Cosmochimica Acta, 31, 2248-2253.

WONES, D.R. (1972) Stability of biotite: A reply. American Mineralogist, 57, 316-317.

ZUREVINSKI, S.E., Heaman, L.M., Creaser, R.A., Strand, P. (2008) The Churchill kimberlite field, Nunavut, Canada: petrography, mineral chemistry, and geochronology. Canadian Journal of Earth Sciences, 45, 1039-1059.

List of Tables

181

List of Tables

Table 2.1. Definition of the 16 J -parameters for phlogopite and their values in eV averaged over runs

for the two compositions x = 0.0 and x = 1.0. The approximate error bar is 0.05 eV. .................. 44

Table 4.1. Crystallite sizes and relative amounts of phases in phlogopite samples determined by

LeBail-fitting of the phlogopite XRD patterns. ............................................................................... 61

Table 4.2. Comparison of relative numbers of Si environments in the tetrahedral sheets of phlogopite.

...................................................................................................................................................... 79

Table 4.3. Comparison of relative numbers of H-OMg3 and H-OMg2Al environments determined from

MC simulations and from 1H MAS NMR spectroscopy. .............................................................. 104

Table 4.4. Initial magnetisation M0, cross-polarisation time THSi, spin-lattice relaxation time in the

rotating frame T1ρ, and mean H-Si distance dH-Si obtained from fits of {1H} → 29Si magnetisation

curves of Al-rich phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y (F)2-y. The error range of THSi is ±

1.0 s, that of dH-Si ± 0.1 Å. ........................................................................................................... 116

Table 4.5. Initial magnetisation M0, cross-polarisation time TFSi, spin-lattice relaxation time in the

rotating frame T1ρ, and mean F-Si distance dF-Si obtained from fits of {19F} → 29Si magnetisation

curves for F-rich phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y (F)2-y. The error range of THSi is ±

1.0 s, that of dH-Si ± 0.1 Å. ........................................................................................................... 118

Table 4.6. Fit parameters obtained from 1D {19F} → 29Si CPMAS NMR spectra of phlogopites K (Mg3-

xAlx) (Al1+xSi3-xO10) (OH)y F2-y recorded with different contact times. For comparison fit parameters

for 29Si MAS NMR spectra are also given. Pos. = position, FWHM = full width at half maximum,

F = relative signal area. The approximate error range for the signal area is ± 2 %. .................. 120

Table 4.7. NMR parameters obtained from 27Al MAS NMR spectra of phlogopite samples

K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)2 recorded at 17.6 T (27Al Larmor frequency = 195.28 MHz). ... 126

Table 4.8. Parameters obtained from 27Al MQMAS NMR spectra of phlogopites and Al2O3, recorded at

16.45 T. ....................................................................................................................................... 131

Table 4.9. Quadrupolar coupling parameters obtained from fits of 17O MAS and MQMAS NMR spectra

obtained at 9.34 T. Error ranges have been estimated by changing the parameters manually

observing χ2 until a disting change of χ2 took place. .................................................................. 138

Table 4.10. Crystallite sizes and relative amounts of phases in phlogopite samples determined by

LeBail-fitting of the phlogopite XRD patterns. The given R-value is that of the overall fit. ......... 140

Table 6.1. List of abbreviations ........................................................................................................... 150

Table 6.2. Parameters obtained from 29Si MAS NMR spectra. ........................................................... 153

Table 6.3. NMR spectroscopic parameters obtained from 1H MAS NMR spectra. ............................. 157

Table 6.5. Parameters obtained from 27Al MAS and MQMAS NMR spectra. ..................................... 162

Table 6.6. {1H} → 29Si contact time dependent magnetisation curves. A = overall intensity of Si-nAl

signals. I = point of highest intensity of the Si-Si2Al or Si-SiAl2 signal. ....................................... 166

Table 6.7. {19F} → 29Si contact time dependent magnetisation curves. A = overall intensity of Si-nAl

signals. I = point of maximum intensity of the Si-Si2Al or Si-SiAl2 signal. ................................... 168

List of Figures

183

List of Figures

Figure 2.1. View on the stacking sequence of phlogopite-2M1. The unit cell is outlined. After Hendricks

and Jefferson (1939). ...................................................................................................................... 9

Figure 2.2. View on the octahedral sheet of phlogopite-2M1 (after Hendricks and Jefferson, 1939). ... 10

Figure 2.3. View on the tetrahedral sheet of phlogopite-2M1 (after Hendricks and Jefferson, 1939). In

the right half of the picture the K+-ions were omitted to show the position of the OH/F site. ....... 11

Figure 2.4. Schematic illustration of different ways of stacking in micas leading to a different position of

the OH-groups. ............................................................................................................................. 12

Figure 2.5. Crystal structures of the five naturally occurring polytypes in micas (Ferraris and Ivaldi,

2002). ............................................................................................................................................ 13

Figure 2.6. Sketch of two rings of tetrahedra belonging to adjacent layer packages. In between, the

interlayer cation K+ is shown. In dioctahedral micas the proton of the OH-group is pointing into

the vacancy, minimizing the repulsion between like-charged proton and K+. (Brigatti and

Guggenheim, 2002) ...................................................................................................................... 17

Figure 2.7. The dipolar interaction between two spins i and j. .............................................................. 24

Figure 2.8. Schematic illustration of the changes of the differences between the energy levels for

Zeeman, first-order and second-order quadrupolar interaction for a spin 3/2 nucleus (after Medek

et al., 1998). .................................................................................................................................. 27

Figure 2.9. Sketch of the energy levels of 1H (‘cold spin revervoir’, left) and 29Si (‘hot spin reservoir’,

right). A transfer of energy from the hot system to the cold one is only possible if the Hartmann-

Hahn-condition is fulfilled (middle). ............................................................................................... 29

Figure 2.10. Pulse sequence schemes for {1H} → 29Si CPMAS (a) and 2D {1H} → 29Si HETCOR (b)

NMR experiments. ........................................................................................................................ 31

Figure 2.11. Example of a magnetisation function for the case of a large proton spin reservoir. The

curve has been calculated according to equation (2.33) using the following parameters: M0 =

4*1010 a.u., T1ρ = 45 ms, and THSi = 9 ms. .................................................................................... 32

Figure 2.12. Example of a magnetisation function for the case of an isolated spin system. The curve

has been calculated according to equation (2.34) using the following parameters: M0 = 4*1010

a.u., T1ρ = 12 s, THSi = 6 ms, THH = 50 ms, a = 0, and b = 500. .................................................... 33

Figure 2.13. Top: Pulse scheme for the 27Al 3QMAS NMR experiment. Bottom: Corresponding

coherence path scheme. .............................................................................................................. 37

Figure 2.14. Assignment of J -parameters within one tetrahedral sheet. ............................................. 41

Figure 2.15 Assignment of octahedral J -parameters. ......................................................................... 42

Figure 2.16. Examples of tetrahedral intralayer (green), tetrahedral interlayer (blue), and octahedral-

tetrahedral (red) J -parameters. ................................................................................................... 43

Figure 4.1. Scanning electron microscope (SEM) images of typical run products. The samples consist

of several µm large crystals of impurity phases (a) with much smaller phlogopite crystals sticking

to them (b). The phlogopite platelets exhibit a diameter of less than 1 µm and often show a more

or less hexagonal shape (c). ......................................................................................................... 60

List of Figures

184

Figure 4.2. Comparison of 29Si MAS NMR spectra of phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y

with different Al- and F-contents. Below the spectra, the Al-content of the tetrahedral sheets

calculated from the relative signal intensities, xest, is given. ......................................................... 65

Figure 4.3. a) The Si/[4]Al-ratio calculated from the relative signal intensities in the 29Si MAS NMR

spectra plotted against the nominal Al-content of the initial oxide mixture. The solid black curve

indicates the phlogopite composition if all starting material had reacted to phlogopite. b) Plot of

the experimentally derived (additional) Al-content of the tetrahedral sheets of the phlogopites

against the Al-content of the initial gel mixture. The black line indicates a complete reaction of the

starting material to phlogopite. ...................................................................................................... 67

Figure 4.4. Comparison of 29Si MAS NMR spectra of phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y

with high Al- and low F-contents. Below the spectra, the Al-content of the tetrahedral sheets

calculated from the relative signal intensities, xest, is given. ......................................................... 69

Figure 4.5. a) The Si/[4]Al-ratio calculated from the relative signal intensities in the 29Si MAS NMR

spectra plotted against the nominal Al-content of the initial oxide mixture. The solid black curve

indicates the phlogopite composition if all starting material had reacted to phlogopite. b) Plot of

the experimentally derived (additional) Al-content of tetrahedral sheets of the phlogopites against

the Al-content of the initial gel mixture. The black line indicates a complete reaction of the

starting material to phlogopite. ...................................................................................................... 70

Figure 4.6. Comparison of 29Si MAS NMR spectra of phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y

with different Al- and F-contents. .................................................................................................. 71

Figure 4.7. Comparison of 29Si MAS NMR spectra of F-rich phlogopites with nominal composition

K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y. ............................................................................................ 73

Figure 4.8. Comparison of the tetrahedral intrasheet J-parameters. The error range of the values is ±

0.05 eV. ......................................................................................................................................... 75

Figure 4.9. Configuration of lowest energy for ordering of cations in a single tetrahedral sheet of

phlogopite with x = 1.0 (‘eastonite’ composition, K (Mg2Al) (Al2Si2O10) (OH)2). Al-atoms are

shown in red, Si-atoms in yellow. Grey bars indicate Al-Si-neighbour pairs. Only a part of the

supercell is shown. Note the defects characterised by Al-Al neighbour pairs. ............................. 75

Figure 4.10.Configurations of lowest energies for cation ordering in a single tetrahedral sheet of

phlogopite with a) x = 0.5 (composition K (Mg2.5Al0.5) (Al1.5Si2.5O10) (OH)2) and b) x = 0.25

(composition K (Mg2.25Al07.5) (Al1.75Si2.25O10) (OH)2). Al-atoms are shown in red, Si-atoms in

yellow. Grey bars indicate Al-Si-neighbour pairs. Only a part of the supercell is shown. ............ 76

Figure 4.11. Configurations of lowest energy for cation ordering in a single tetrahedral sheet of

phlogopite with x = 0.0 (composition K Mg3 (AlSi3O10) (OH)2). Al-atoms are shown in red, Si-

atoms in yellow. Every Al-atom has three Si-atoms as next-nearest-neighbours, while every Si-

atom is surrounded by two Si-atoms and one Al-atom in the neighbouring tetrahedra. a) Grey

lines indicate Si-Al neighbour pairs. b) The J3-interactions connecting Al-atoms are marked by

grey lines. ...................................................................................................................................... 78

Figure 4.12. Comparison of experimental 29Si MAS NMR spectra (right) and theoretical ones derived

from the Monte-Carlo simulation results (left) for phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)2.

List of Figures

185

For x = 0.0, the 29Si MAS NMR spectrum of a F-bearing phlogopite is shown as no F-free

samples have been available for analysis. ................................................................................... 80

Figure 4.13. 1H MAS NMR spectra of phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y with different

OH- and Al-contents. Below the spectra the ratio I[H-OMg2Al]/(I[H-OMg2Al] + I[H-OMg3]) is given,

abbreviated as ‘Al/(Mg+Al)’. .......................................................................................................... 83

Figure 4.14. Plot of the relative signal intensity of the H-OMg2Al signal against the Al-content x

estimated from 29Si MAS NMR spectra. The solid line represents a statistical distribution of ions

in the octahedral sheet. a) Data of this study only, b) comparison of the 600 °C data (black

symbols) to the 800 °C data of Fechtelkord et al. (2003a, grey symbols). ................................... 84

Figure 4.15. Comparison of 19F MAS NMR spectra of phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y

with different Al- and F-contents. Below the spectra the ratio I[F-Mg2Al]/(I[F-Mg2Al] + I[F-Mg3]) is

given, abbreviated as ‘Al/(Mg+Al)’. Spinning sidebands are marked by asterisks. ...................... 87

Figure 4.16. Plot of the relative intensity of the F-OMg2Al signal against the Al-content x estimated

from 29Si MAS NMR spectra. The solid line represents a statistical distribution of ions in the

octahedral sheet. .......................................................................................................................... 88

Figure 4.17. Comparison of 1H MAS NMR spectra of OH- and Al-rich phlogopites

K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y. Below the spectra the ratio

I[H-OMg2Al]/(I[H-OMg2Al] + I[H-OMg3]) is given, abbreviated as ‘Al/(Mg+Al)’. ............................ 90

Figure 4.18. Plot of the relative signal intensity of the H-OMg2Al signal against the Al-content xest

estimated from 29Si MAS NMR spectra. The solid line represents a statistical distribution of ions

in the octahedral sheet. ................................................................................................................ 92

Figure 4.19. Plot of the H-OMg2Al (a) and the H-OMg3 (b) signal position as a function of the Al-

content of the estimated Al-content of the phlogopites. Tolerances have been estimated by

changing parameters manually observing χ2 until a distinct change of χ2 took place. ................. 92

Figure 4.20. Comparison of 19F MAS NMR spectra of OH- and Al-rich phlogopites

K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y. Below the spectra the ratio I[F-Mg2Al]/(I[F-Mg2Al] + I[F-Mg3])

is given, abbreviated as ‘Al/(Mg+Al)’. Spinning sidebands are marked by asterisks. .................. 94

Figure 4.21. Position and full width at half maximum (FWHM) of 19F MAS NMR signals versus

estimated Al-content of phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y. a)+c) F-Mg3 signal.

b)+d) F-Mg2Al signal. Tolerances have been estimated by changing parameters manually

observing χ2 until a distinct change of χ2 took place. .................................................................... 95

Figure 4.22. 19F MAS NMR spectrum of sample with nominal composition of xnom = 1.2 and y = 1.6

showing a splitting of the F-Mg3 signal at -175 ppm into two separate signals. Spinning

sidebands are marked by asterisks. ............................................................................................. 96

Figure 4.23. Comparison of 1H MAS NMR spectra of phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y.

Below the spectra the ratio I[H-OMg2Al]/(I[H-OMg2Al] + I[H-OMg3]) is given, abbreviated as

‘Al/(Mg+Al)’.................................................................................................................................... 98

Figure 4.24. Comparison of 19F MAS NMR spectra of F-rich phlogopites

K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y. Below the spectra the ratio I[F-Mg2Al]/(I[F-Mg2Al] + I[F-Mg3])

is given, abbreviated as ‘Al/(Mg+Al)’. Spinning sidebands are marked by asterisks. .................. 99

List of Figures

186

Figure 4.25. Comparison of 19F MAS NMR spectra of phlogopites of composition xnom = 0.8, y = 0.5

synthesised at 600 (a) and at 800 °C. ........................................................................................ 100

Figure 4.26. Comparison of the octahedral intrasheet J-parameters. The error range of the values is ±

0.05 eV. ....................................................................................................................................... 101

Figure 4.27. Configuration of lowest energy derived by Monte-Carlo simulations for a single octahedral

sheet of phlogopites with composition a) x = 0.25, y = 2.0 (K (Mg2.75Al0.25) (Al1.25Si2.75O10) (OH)2)

and b) x = 0.75, y = 2.0 (K (Mg2.25Al0.75) (Al1.75Si2.25O10) (OH)2). Mg-ions are shown in green, Al-

atoms in red. Grey bars indicate Mg-Al neighbour pairs. Only a part of the supercell is shown. 102

Figure 4.28. Configuration of lowest energy derived by Monte-Carlo simulations for a single octahedral

sheet of phlogopite with composition x = 1.0, y = 2.0 (K (Mg2Al) (Al2Si2O10) (OH)2). Mg-ions are

shown in green, Al-atoms in red. Grey bars indicate Mg-Mg neighbour pairs. Only a part of the

supercell is shown. ...................................................................................................................... 103

Figure 4.29. 2D {1H} → 29Si HETCOR CPMAS NMR spectrum of phlogopite with nominal composition

K(Mg2.2Al1.8)(Al1.8Si2.2O10)(OH)2 (xnom = 0.5)................................................................................ 107

Figure 4.30. View on the tetrahedral sheets of phlogopite. Every OH-position is co-ordinated by three

octahedral cations which may be either Mg or Al (white arrows). The information on a single OH

environment is passed on to six neighbouring tetrahedra if these are occupied by 29Si (black

arrows). Each tetrahedral site has three next-nearest-neighbours which may be either Si or Al. In

this way, the number of Al co-ordinating OH may be correlated to the amount of Al in the 29Si

environment in the tetrahedral sheet. ......................................................................................... 107

Figure 4.31. Values of the tetrahedral intralayer, tetrahedral interlayer, and the octahedral –

tetrahedral interaction parameters obtained from GULP. ........................................................... 109

Figure 4.32. Details of the configuration of lowest energy obtained from MC simulations for x = 1.0.

Only one 2:1 layer package is given. Si, [4]Al, Mg, and [6]Al-atoms are shown in yellow, red, blue,

and green, respectively. In the lower picture, Mg has been omitted for clarity. Grey bars connect

pairs of directly neighboured Si- and Al-atoms. Domains can be distinguished by the different

orientation of [4]Al in the lower tetrahedral sheet to [4]Al in the upper tetrahedral sheet. Two such

configurations are marked by white ellipsoids. Some of the domain boundaries are highlighted by

white lines. They are characterised by Si-O-Si and Al-O-Al linkages. ........................................ 110

Figure 4.33. Details of the configuration of lowest energy obtained from MC simulations for x = 0.5.

Only one 2:1 layer package is given. Si, [4]Al, Mg, and [6]Al-atoms are shown in blue, red, green,

and yellow, respectively. In the lower picture, Mg has been omitted for clarity. Grey bars indicate

Al-O-Si linkages in the tetrahedral sheet. ................................................................................... 112

Figure 4.34. Comparison of site connectivies obtained from MC simulations for a composition of x =

0.68 (left) to the 2D {1H} → 29Si CPMAS HETCOR NMR spectrum of a phlogopite with the same

estimated Al-content. .................................................................................................................. 113

Figure 4.35. Experimental magnetisation curves for signal area (top, all Si-nAl signals) and highest

signal intensity (bottom, Si-Si2Al or Si-SiAl2 signal). ................................................................... 115

Figure 4.36. Magnetisation curve derived from the highest intensities of the Si-Si2Al signal for

phlogopite of nominal composion xnom = 0.7, y = 1.0 (synthesis temperature T = 800 °C). The

solid line represents a fit to the data according to equation (2.33). ............................................ 116

List of Figures

187

Figure 4.37. Experimental magnetisation curves of F-rich phlogopites

K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y derived from the signal area of all Si-nAl signals. Top: Whole

contact time range. Bottom: Detail of low contact times. ............................................................ 117

Figure 4.38. Comparison of 27Al MAS NMR spectra of phlogopites of different compositions

K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y. Spinning sidebands are marked by asterisks. ................. 123

Figure 4.39. Plot of the phlogopite [4]Al signal position against the estimated Al-content xest of

phlogopites with different F-contents y. ...................................................................................... 124

Figure 4.40. 27Al MQMAS NMR spectra of phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y. Spinning

sidebands are marked by asterisks. The F1-axis has been labelled according to the C3a-

convention (Amoureux and Fernandez, 1998; Millot and Man, 2002) ........................................ 125

Figure 4.41. Comparison of 27Al MAS NMR spectra of phlogopite samples with nominal composition

K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y recorded at 17.6 T. ........................................................... 127

Figure 4.42. Comparison of 27Al MAS NMR spectra of phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)2 F2

and an Al-oxide sample recorded at a field strength of 16.45 T (27Al Larmor frequency

182.47 MHz). .............................................................................................................................. 129

Figure 4.43. 27Al MAS and MQMAS spectra of Al2O3. The 27Al MAS NMR spectrum is shown on top of

the F2-projection of the 27Al MQMAS spectrum. In the left part slices parallel to the F2-axis of the

MQMAS spectrum are shown, and the F1-shifts at which they were taken are given. Labelling of

the F1-axis has been done following the Cz-convention (Millot and Man, 2002) The diagonal line

in the MQMAS spectrum indicates positions resulting from Al environments of high symmetry. In

these, no electric field gradient influences the nucleus and thus the signal shift is only made up

by the chemical shift. .................................................................................................................. 130

Figure 4.44. 27Al MQMAS NMR spectra of phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y recorded

at 16.45 T (27Al Larmor frequency 182.42 MHz). The F1-axis has been labelled according to the

Cz-convention (Millot and Man, 2002). ........................................................................................ 132

Figure 4.45. 17O MAS NMR spectra of 17O enriched phlogopites K (Mg3-xAlx) (Al1+xSi3-xO10) (OH)y F2-y.

.................................................................................................................................................... 134

Figure 4.46. 17O MQMAS spectra of phlogopite with composition xnom = 0.5, y = 1.0. a) Spectrum

recorded at 9.34 T (F1-axis labelled according to the C3a-convention; Amoureux and Fernandez,

1998; Millot and Man, 2002). b) Spectrum recorded at 16.45 T. ................................................ 135

Figure 4.47. 17O MAS NMR spectrum of phlogopite with composition xnom = 0.5, y=1.0. Observed

spectrum, total lineshape fit, and individual signal components are shown. .............................. 137

Figure 4.48. X-ray diffraction powder patterns of four selected phlogopite samples. Arrows mark peak

positions of the impurity phase corundum. ................................................................................. 139

Figure 4.49. Results of the analysis of XRD powder patterns of several 1M-phlogopites with varying

compositions K(Mg3-xAlx)(Al1+xSi3-xO10)(OH)yF2-y. ....................................................................... 141

Figure 4.50. Sketch showing the distortion of a tetrahedral sheet. a) Undisturbed sheet with hexagonal

symmetry. b) Rotation of tetrahedra about the perpendicular to the sheet leads to a ditrigonal

symmetry. The distortion is described by the ditrigonal rotation angle α. c) Fully distorted

tetrahedral sheet. (Ferraris and Ivaldi, 2002) ............................................................................. 142

List of Figures

188

Figure 4.51. Results of the analysis of XRD powder patterns of several 2M1-phlogopites with varying

compositions K(Mg3-xAlx)(Al1+xSi3-xO10)(OH)yF2-y......................................................................... 143

Figure 4.52. The relative amount of phlogopite-1M of phlogopites derived from LeBail-fitting plotted

against the estimated Al-content of the phlogopites. .................................................................. 143

Figure 4.53. Sketch of the interlayer boundary in phlogopite-1M (left) and phlogopite 2M1 (right).

Tetrahedral tilting, i.e. out-of-plane rotation, is exaggerated. Circles denote K+-ions. Modified

after Ferraris and Ivaldi, 2002. .................................................................................................... 144

Figure 4.54. XRD pattern of phlogopite with nominal composition xnom = 0.4, y = 1.8. a) Whole pattern.

b) Detail. The pattern shows a high background, and the peaks between 20 and 33 °2θ are

surrounded by satellite peaks (marked by arrows) resulting from stacking faults in the structure.

.................................................................................................................................................... 145

Figure 4.55. Results of the analysis of XRD powder patterns of several 2M1-phlogopites with varying

compositions K(Mg3-xAlx)(Al1+xSi3-xO10)(OH)yF2-y......................................................................... 146

Danksagung

189

Danksagung

Zunächst einmal möchte ich all den vielen Leuten danken, die mir in all den Jahren geholfen

haben und ohne die die vorliegende Arbeit nicht möglich gewesen wäre.

Mein ganz besonderer Dank gilt meinem Betreuer Michael Fechtelkord, der mir die Möglichkeit

gab, in einem internationalen Projekt an einem interessanten Thema zu arbeiten. Er stand mir stets

hilfreich zur Seite, war geduldig mit mir und hatte Zeit für zahlreiche Diskussionen, wenn ich mal nicht

weiterkam. Bedanken möchte ich mich auch für die viele Mühe und die Zeit, die er investierte, um mir

ein optimales Arbeitsumfeld zu bieten, beispielsweise durch das stundenlange Werkeln an der

Technik, wenn Hannelore mal wieder streikte, oder die Computeradministration.

Weiterhin danke ich Alberto García Arribas für die Durchführung der Simulationen und für die

Organisation des Projekts „ORION“ sowie der Projekttreffen. In langen, angeregten Diskussionen

gelang es ihm ein ums andere Mal, mir einen anderen Blickwinkel auf das Thema zu eröffnen, Fehler

zu erkennen und noch mehr aus den Daten herauszuholen. Außerdem danke ich ihm auch für die

Übernahme der Zweitkorrektur.

Auch alle anderen Mitglieder des „ORION“-Projektes unterstützten mich stets und trugen durch

Tipps und Anregungen ihren Teil zum Gelingen dieser Arbeit bei.

Predrag Vulić und Volker Kahlenberg synthetisierten einige der in dieser Arbeit untersuchten

Phlogopit-Proben. Tonči Balić-Žunić, Helene Almind und Emil Makovicky halfen bei der Aufnahme und

der Auswertung von Pulverdiffraktogrammen an der Universität Kopenhagen. Karen Friese und

Andrzej Grzechnik untersuchten einige der Proben am DESY und unterstützten mich in Bilbao bei der

Auswertung der Daten. Javier Lopéz-Solano und Lars Olsen litten mit mir bei Vorträgen und

Posterpräsentationen und halfen mir, mich in Bilbao bzw. Kopenhagen zurechtzufinden. Louise

Nielsen danke ich für ihre Bemühungen, möglichst eisenfreie natürliche Phlogopite zu finden.

Besonders möchte ich auch den Mitgliedern des Bereichs Mineralogie-Kristallographie sowie des

Bereichs Mineralogie-Petrologie des Instituts für Geologie, Mineralogie und Geophysik der Ruhr-

Universität Bochum danken.

Antje Grünewald-Lüke sorgte dafür, dass ich nicht ganz so allein war in unserem Büro und sorgte

dafür, dass neben all der Arbeit auch der Spaß nicht zu kurz kam. Zusammen mit Kirsten Keppler,

Tomasz Goral, Sandra Grabowski und Ute Gundert unterstützte sie mich bei den Arbeiten im Labor.

Dank auch an Bernd Marler, der Röntgenmessungen an meinen Proben – in Bochum und am

DESY – vornahm und auch viel Zeit damit zubrachte mir bei der Auswertung der Diffraktogramme zu

helfen, sowie an Thomas Fockenberg für seine Einweisung und Hilfe im Hydrothermallabor.

Weiterhin danke ich Wilfried Schrimpf ganz herzlich für den Einsatz, den er beim schnellen

Reparieren der Probenköpfe und anderem zeigte. Auch Hans-Jochen Hauswald investierte viel Zeit

Danksagung

190

und Mühe bei kleineren und größeren Reparaturen an Hannelore. Benjamin Kellert war stets zur

Stelle, wenn es Probleme im Hydrothermallabor gab. Rolf Neuser nahm mit mir die

elektronenmikroskopischen Bilder auf, und Frank Bettenstedt stand mir beim Schweißen der Kapseln

mit Rat und Tat zur Seite.

Dank gebührt auch den Mitarbeitern der Bruker Biospin GmbH, und hier vor allem Walter Knöller,

die uns an ihrer umfänglichen Erfahrung teilhaben ließen, und es uns durch schnelle und

unkomplizierte Bereitstellung von Leihteilen ermöglichten, zügig Schäden in Hannelores System

ausfindig zu machen und zu reparieren.

Malte Seipenbusch, Lena Lingner, Mareike Wolf, Melanie Lhys-Aliu und Nathalie Lübke möchte

ich dafür danken, dass sie mir als studentische Hilfskräfte einiges an Arbeit abgenommen haben.

Anna Weiner war mir gerade in den letzten Monaten des Zusammenschreibens eine angenehme

Zimmergenossin und half z.B. durch Tipps zum Layout.

Bei Ulrike Werner-Zwanziger und Josef W. Zwanziger bedanke ich mich für die Aufnahme von

Hochfeld-NMR-Spektren in Halifax. Ebenso danke ich der Arbeitsgruppe um Jürgen Haase für weitere

Hochfeld-Messungen in Leipzig.

Last but not least möchte ich mich natürlich besonders bei all jenen bedanken, die mich immer

unterstützten, zu mir hielten und es mir nicht übel nahmen, wenn die Arbeit das ein oder andere Mal

vorging, die mich immer wieder „zwangen“ auch das Leben neben der Arbeit nicht zu kurz kommen zu

lassen, die mich aber andererseits auch immer wieder anspornten und neu motivierten: mein Freund

Jan, meine Familie - Mutsch, Wolfgang, Chris und Max -, Bianca, Eva, Martin und Sonja.

Diese Arbeit wurde mit Mitteln der European Science Foundation (ESF) und der Deutschen

Forschungsgemeinschaft (DFG) gefördert.

Lebenslauf

191

Lebenslauf

Persönliche Daten:

Name Ramona Langner

Geburtsdatum 10.02.1982

Geburtsort Rochlitz

Familienstand ledig

Staatsangehörigkeit deutsch

Schulausbildung

1988 – 1992 Diesterweg-Grundschule Geringswalde

1992 – 1995 Martin-Luther-Gymnasium Hartha

1995 – 2001 Gymnasium Neckargemünd

06 / 2001 Abschluss mit Abitur

Studium

10 / 2001 – 03 / 2002 Grundstudium der Chemie an der Ruprecht-Karls-

Universität Heidelberg

04 / 2002 – 09 / 2003 Grundstudium der Mineralogie an der Ruprecht-

Karls-Universität Heidelberg

01 / 2004 Diplomvorprüfung Mineralogie

10 / 2003 – 01 / 2007 Hauptstudium der Mineralogie an der Ruprecht-

Karls-Universität Heidelberg

Schwerpunkt Kristallographie

Lebenslauf

192

26.01.2007 Diplomprüfung Mineralogie

Thema der Diplomarbeit:

Wachstumstexturen in synthetischen

Eisentitanoxidproben

Betreuer:

Prof. Dr. D. Lattard

seit 04 / 2007 Promotionsstudium an der Ruhr-Universität

Bochum

Tätigkeiten

08 / 2003 – 12 / 2006 Wissenschaftliche Hilfskraft am Institut für

Mineralogie der Ruprecht-Karls-Universität

Heidelberg

02 / 2007 – 01 / 2010 Wissenschaftliche Angestellte am Institut für

Geologie, Mineralogie und Geophysik der

Ruhr-Universität Bochum

Erklärung

193

Erklärung

Hiermit versichere ich an Eides statt, dass ich die vorgelegte Dissertation selbst

verfasst und mich keiner anderen als der von mir ausdrücklich bezeichneten Quellen

und Hilfen bedient habe.

Ich erkläre hiermit, dass ich an keiner anderen Stelle ein Prüfungsverfahren

beantragt bzw. die Dissertation in dieser oder anderer Form bereits anderweitig als

Prüfungsarbeit verwendet oder einer anderen Fakultät als Dissertation vorgelegt

habe.

Bochum, den 14. Juli 2010

Ramona Langner