You are on page 1of 273

Characterization of Local Structures in Layered Niobates

Using Solid-State NMR and Rapid Microwave-Assisted


Synthesis and Exfoliation of Their Alcohol Grafted
Derivatives

JOSHUA ROSS BOYKIN

DECEMBER 2017

A DISSERTATION

Submitted to the faculty of Clark University,


Worcester, Massachusetts,
in partial fulfillment of
the requirements for
the degree of Doctor of Philosophy
in the Department of Chemistry




ProQuest Number: 10683811




All rights reserved

INFORMATION TO ALL USERS
The quality of this reproduction is dependent upon the quality of the copy submitted.

In the unlikely event that the author did not send a complete manuscript
and there are missing pages, these will be noted. Also, if material had to be removed,
a note will indicate the deletion.






ProQuest 10683811

Published by ProQuest LLC (2018 ). Copyright of the Dissertation is held by the Author.


All rights reserved.
This work is protected against unauthorized copying under Title 17, United States Code
Microform Edition © ProQuest LLC.


ProQuest LLC.
789 East Eisenhower Parkway
P.O. Box 1346
Ann Arbor, MI 48106 - 1346
DISSERTATION COMMITTEE

Luis Smith, Ph.D.


Chief Instructor

Mark Turnbull, Ph.D.


Committee Member

Noel Lazo, Ph.D.


Committee Member
ABSTRACT
The research described in this dissertation spans nearly six years and can be

broken into three heavily related research projects:

1. Exploration of local structural changes in mixed alkali layered perovskites.

2. Determination of B-site preference in mixed Nb/Ta systems using solid-state

NMR

3. Microwave assisted grafting and exfoliation of layered perovskites using

long chain alcohols.

Exploration of local structural changes in mixed alkali layered perovskites.

A series of three-layered Dion-Jacobson (D-J) RbSrxCa2-xNb3O10 (0 < x < 2)

and RbBaxSr2-xNb3O10 (0 < x < 0.8) compounds were synthesized using microwave

and molten salt synthetic routes, respectively, and analyzed using powder X-Ray

diffraction (XRD), scanning electron microscopy (SEM), energy dispersive

spectroscopy (EDS), attenuated total reflectance-infrared spectroscopy (ATR-IR),

and wideline uniform rate smooth truncation-quadrupolar Carr-Purcell-Meiboom-

Gill (WURST-QCPMG) nuclear magnetic resonance (NMR). An unexpected

dichotomy regarding cation distribution was observed with distribution considered

homogeneous on a long-range scale and heterogeneous on a local scale.

Determination of B-site preference in mixed Nb/Ta systems using solid-state NMR

D-J phase layered perovskites can readily undergo proton exchange under

mild conditions leading to the formation of a solid acid capable of intercalating


various organic molecules. The B-site atom in this class of compounds is often

Ti4+, Nb5+, or Ta5+. It has been shown that changing the B-site atom can

drastically alter the intercalation behavior and catalytic ability of these materials.

While Nb5+ and Ta5+ possess the same charge and approximately equal ionic radii

(0.64 Ǻ), Nb is more electronegative than Ta (1.6 vs. 1.5 Pauling Units,

respectively) leading to Nb-O bonds being more covalent in character than Ta-O

bonds and changes in the energy states of Ta5+ and Nb5+ compounds. As this

work shows, in mixed Nb5+/Ta5+ systems, electronegativity is a strong driver of

site ordering in D-J niobates.

Microwave assisted grafting and exfoliation of layered perovskites using long chain
alcohols.
Current grafting methods of layered perovskites involve high pressure

heating at 80 to 150 oC for days to weeks, depending on the alcohol involved. The

long reaction time involved inhibits the ability to study grafted samples in a timely

fashion. This work will present a new microwave assisted grafting technique which

decreases reaction time from 17 days to 12 hours. Furthermore, a microwave-

sonication based method of exfoliation of grafted niobates in organic solvents was

developed, further expanding the capabilities and potential of niobate nanosheets.


© 2017

JOSHUA ROSS BOYKIN

ALL RIGHTS RESERVED


Academic History

Name (in Full): Joshua Ross Boykin Date: AUG 2017

Baccalaureate Degree: Bachelor of Science (Chemistry)

Source: Marlboro College, Marlboro, VT Date: May, 2008

Occupation and Academic Connection since date of baccalaureate


degree:
Organic Chemistry Laboratory Instructor, Clark University: 2016-2017
Introductory Chemistry Instructor, Clark University: 2016
Teaching Assistant, Clark University: 2010-2016
DEDICATION

To the steadfast rejection of ignoramus et ignorabimus and all it


embodies.

Also, to my friends and family who will never read this dissertation, but
will still tell me it’s great.

vii
ACKNOWLEDGEMENTS

I have spent the past seven years completing my Ph.D. at Clark University.
I would like to extend my sincerest gratitude to everyone in the chemistry
department for helping me to reach this point. There were many times I became
unsure of myself and seriously doubted whether I could accomplish this feat, and
it was only through the strength of those around me that I find myself at the end
of this long journey. To name every person and every act that has lead me here
would result in a tome that would likely rival in length this dissertation, but I would
like to take a little time to give credit to a few.

First, I would like to thank my advisor Professor Luis Smith. When I started
at Clark I knew nothing beyond the very basics of solid-state NMR, X-ray
diffraction, or inorganic materials. He has shown great patience as I explored the
worlds of spectroscopy and perovskites, giving me guidance, encouragement, and
when need be criticism. He has served as an excellent example during my time
at Clark and I thank him for it. He has been the best mentor I could have asked
for.

Second, I would like to thank my other committee members, Professors


Mark Turnbull and Noel Lazo. During my very first semester at Clark I took courses
from both and it was through them that I came to realize the hard work that would
be required of me during graduate school. Both have helped me greatly since
then by asking me difficult questions about my work that have forced me to
consider different perspectives that I would not have otherwise considered.

Thirdly, I would like to thank Dr. Lin, Clark’s NMR manager. Reading about
new pulse sequences or doing theoretical calculations don’t mean much if you
can’t operate the instrument. He has helped me learn the hardware and software
considerations that are not discussed, but nonetheless necessary for success.

Last, but by no stretch of the imagination least, I would like to thank my


family for their love and support. Since I was a child my parents always
encouraged me in my desire to be a scientist and at my lowest points were there
to help me rise. My lovely wife, Jodi, has made me happier than I ever thought I
could be, and I know without a shadow of a doubt that I wouldn’t be here without
her. To those I have mentioned, and the many more unmentioned, I am infinitely
indebted.

viii
Table of Contents
List of Tables ............................................................................................................................. xii
List of Figures ........................................................................................................................... xv
List of Abbreviations ........................................................................................................... xxxi
Chapter 1: Background of Layered Perovskites ......................................................................... 1
1.1 Introduction ............................................................................................................................ 1
1.2 RbCaxSr2-xNb3O10 and RbCa2NaNb4-xTaxO13 Crystal Structures ........................................... 6
1.3 Soft-chemical Exchange Reactions in Layered Perovskites .................................................. 9
1.4 Niobate Nanosheets ............................................................................................................. 11
1.5 Research Presented Herein .................................................................................................. 14
1.5.1 Exploration of local structural changes in mixed alkali layered perovskites. ................... 14
1.5.2 Determination of B-site preference in mixed Nb/Ta systems using solid-state NMR ...... 15
1.5.3 Microwave assisted grafting and exfoliation of layered perovskites using long chain
alcohols. ..................................................................................................................................... 16
Chapter 2: NMR Background .................................................................................................... 18
2.1 Solid-State NMR.................................................................................................................. 18
2.2 Quadrupolar Nuclei .............................................................................................................. 20
2.3 93Nb Quadrupolar Interactions ............................................................................................. 30
2.4 Variable Offset Cumulative Spectra (VOCS) ...................................................................... 31
2.5 Magic Angle Spinning ......................................................................................................... 33
2.6 Wideline Uniform Rate Smooth Truncations (WURST) Pulses.......................................... 36
2.7 Multiple-Quantum Magic Angle Spinning (MQMAS) ....................................................... 42
Chapter 3: Experimental Overview ........................................................................................... 49
3.1 Materials .............................................................................................................................. 49
3.2 Synthetic Procedures*.......................................................................................................... 49
3.2.1 Three-layer Niobates Using MSS .................................................................................. 49
3.2.2 Three-layer Niobates Using Microwave Heating ......................................................... 50
3.2.3 Alkyl Grafting with Conventional Heating ................................................................... 50
3.2.4 Alkyl Grafting with Microwave Heating ....................................................................... 51
3.2.5 RbCa2NaNb4-xTaxO13 and HCa2NaNb4-xTaxO13 ............................................................. 52
3.2.6 Nanosheet synthesis using TBAOH ............................................................................... 53
3.2.7 Nanosheet Synthesis Using Microwave Assisted Grafting............................................ 53
3.3 Characterization Methods .................................................................................................... 54

ix
3.3.1 Powder XRD ................................................................................................................. 54
3.3.2 ATR-IR .......................................................................................................................... 54
3.3.3 SEM and EDX ............................................................................................................... 55
3.3.4 UV-visible Diffuse Reflectance ..................................................................................... 55
3.3.5 Solid-State NMR Experiments ....................................................................................... 55
Chapter 4. Exploration of Local Structural Changes in Layered Niobates due to
Compositional Changes at the A-Site. ........................................................................................ 58
4.1 Introduction .......................................................................................................................... 58
4.2 Experimental ........................................................................................................................ 59
4.3 XRD Results and Discussion ............................................................................................... 61
4.4 EDS and SEM Results and Discussion ................................................................................ 72
4.5 ATR-IR Results and Discussion .......................................................................................... 75
4.6 NMR Results and Discussion .............................................................................................. 77
4.7 Conclusions ........................................................................................................................ 109
Chapter 5: Determination of Cation Site Ordering in Four-layer HCa2NaNb4-xTaxO13
Compounds Using Solid State 93Nb and 23Na NMR ................................................................ 111
5.1 Introduction ........................................................................................................................ 111
5.3 Results and Discussion ...................................................................................................... 115
5.3.1 XRD, SEM, and Elemental Analysis ........................................................................... 115
5.3.2 93Nb NMR .................................................................................................................... 119
5.3.3 23Na MAS NMR ........................................................................................................... 131
5.4 Conclusions ........................................................................................................................ 135
Chapter 6: Microwave Assisted Grafting of Mixed Cation HCaxSr2-xNb3O10
Compounds with n-Alcohols* ................................................................................................... 136
6.1 Introduction ........................................................................................................................ 136
6.2 Experimental ...................................................................................................................... 137
6.3 XRD Results and Discussion ............................................................................................. 139
6.4 TGA and 1H NMR Results and Discussion ....................................................................... 145
6.5 Grafting of Mixed Ca/Sr Compounds ................................................................................ 147
6.6 UV-visible Results and Discussion .................................................................................... 152
6.7 Conclusions ........................................................................................................................ 155
Chapter 7: Exfoliation of Grafted Niobates Using Organic Solvents................................... 156
7.1 Introduction ........................................................................................................................ 156

x
7.2 Experimental ...................................................................................................................... 158
7.2.1 Sample Overview ........................................................................................................ 158
7.2.2 Solvent Choice ............................................................................................................ 161
7.2.4 Sonication ................................................................................................................... 162
7.2.5 Sedimentation.............................................................................................................. 162
7.2.6 Extracting Nanosheets Out of Solution ....................................................................... 163
7.3 Visual Observations and Generation of Colored Suspensions........................................... 163
7.4 XRD Results ...................................................................................................................... 173
7.2.3 Heating........................................................................................................................ 174
7.4.1 Sediments of Suspensions ............................................................................................ 175
7.4.2 Dried Suspensions ....................................................................................................... 178
7.4.3 Solid From SN ............................................................................................................. 183
7.5 SEM and EDS Results ....................................................................................................... 187
7.6 TGA Results and Exfoliation Efficiency ........................................................................... 193
7.7 Discussion .......................................................................................................................... 194
7.7.1 Role of Solvent and Parent Compound ....................................................................... 194
7.7.2 Role of Sonication ....................................................................................................... 195
7.7.3 Role of Microwave Heating ........................................................................................ 195
7.7.4 Treatment of Supernatant ........................................................................................... 196
7.7.5 Fate of the Grafted Alkyl Groups ................................................................................ 196
7.8 Conclusions ........................................................................................................................ 198
Chapter 8: Summary and Future Directions .......................................................................... 200
References ................................................................................................................................... 204
Appendix A: Indexed Lists of XRD Pattern hkl Reflections.................................................. 219
Appendix B: C1-6/CaxSr2-xNb3O10 TGA Data ......................................................................... 233

xi
List of Tables
Table 1.1. Tetragonal unit cell parameters of RbCaxSr2-xNb3O10 phases collected by
Geselbracht et al.31 ................................................................................................... 7

Table 1.2 Nb-O bond lengths in RbSr2Nb3O10. .......................................................................... 8

Table 3.1. General grafting procedure for CaxSr2-xNb3O10 compounds. .................................. 52

Table 4.1. Synthetic conditions for RbCaxSr2-xNb3O10 compounds using dry microwave
heating or MSS. ..................................................................................................... 60

Table 4.2. Synthetic conditions for RbBaxSr2-xNb3O10 compounds using dry microwave
synthesis................................................................................................................. 61

Table 4.3. Lattice parameters for RbCaxSr2-xNb3O10 compounds obtained from Le Bail
fitting using TOPAS software. .............................................................................. 63

Table 4.4. Lattice constants for RbBaxSr2-xNb3O10 compounds............................................... 63

Table 4.5. Atom percent results for RbCaxSr2-xNb3O10 obtained using EDS. Ca content was
estimated using ratio of Nb:Ca due to clean separation of peaks in EDS spectra.
Ca content error was calculated using Fieller’s method of estimated confidence
intervals between ratios. ........................................................................................ 72

Table 4.6. Atom percent results for RbBaxSr2-xNb3O10 obtained using EDS. Ba content was
estimated using ratio of Nb:Ca due to clean separation of peaks in EDS spectra.
Ba content error was calculated using Fieller’s method of estimated confidence
intervals between ratios. ........................................................................................ 73

Table 4.7. First moment analysis for RbCaxSr2-xNb3O10 compounds. The obtained CQ and
η values were calculated using Equation 2.42-45. ................................................. 89

Table 4.9. RbCaxSr2-xNb3O10 NMR Fit parameters. Errors in the last digit are shown in
parentheses and were calculated for population and CQ values using chi-square
minimization. ......................................................................................................... 93

Table 4.10. RbBaxSr2-xNb3O10 93Nb Static NMR Fit parameters. Errors in the last digit are
shown in parentheses and were calculated for population and CQ values using
chi-square minimization. ..................................................................................... 103

xii
Table 5.1. Le Bail method fit lattice parameters, particle size, zero correction, χ2, and Rwp*
for XRD patterns shown in Figures 3 and 4 ........................................................ 119

Table 5.2. NMR fit parameters for RbCa2NaNb4O13 and RbCa2NaNb2.8Ta1.2O13 obtained
from VOCS spectra. Error in the last digit is shown in parentheses................... 124

Table 5.3. 93Nb NMR fit parameters for HCa2NaNb4O13 and HCa2NaNb2.8Ta1.2O13 obtained
from MQMAS and VOCS spectra. Error in the last digit is shown in
parentheses........................................................................................................... 130

Table 5.4 93Nb NMR fit parameters for HCa2NaNb4O13 and HCa2NaNb2.8Ta1.2O13 BRAIN-
CP spectra. ........................................................................................................... 130

Table 5.5. 23Na NMR fit parameters obtained from MQMAS and MAS spectra. ................. 134

Table 5.6 Results of EFG tensor calculations using a point charge model for
HCa2NaNb4O13. ................................................................................................... 135

Table 6.1. Heating Cycles and Reaction Times for Microwave and Conventional Heating
Methods ............................................................................................................... 139

Table 6.2. Le Bail method fit lattice parameters, zero correction, χ2, particle size in stacking
direction, and Rwpa from Le Bail fits for hydrated and C6 samples. Error in the
last digit is represented by parentheses. ............................................................... 144

Table 6.3. Alkyl coverage percentages for C1-C6/CaxSr2-xNb3O10 compounds using


microwave grafting. ............................................................................................. 148

Table 7.1 Summary of exfoliation procedure by sample. More detailed descriptions of the
exfoliation procedures are also in the text. .......................................................... 159

Table 7.1 Cont’d. ................................................................................................................... 160

Table 7.2 Hansen Solubility Parameters for exfoliation solvents and alcohols used during
grafting of parent compounds. Values are given in units of MPa1/2. Values
taken from Hansen.197 .......................................................................................... 161

Table 7.3. XRD Stacking Reflections Summary ................................................................... 174

Table 7.4. Atomic composition of samples 14C and 15C using copper tape as substrate.
The letter in parentheses indicates to which image in Figure 7.28 above the data

xiii
correspond. Error was automatically calculated by the Quantax EDS
application. .......................................................................................................... 191

xiv
List of Figures
Figure 1.1. Schematic representation of the ideal perovskite structure based on the cubic
SrTiO3 crystal structure. Left- Ball and stick model of unit cell where blue
spheres = Ti, red spheres = O, and green sphere = Sr. Right- polyhedral model
showing TiO6 octahedra in blue and Sr occupying interstitial 12-coordinate
sites in green. All crystallographic representations have been generated using
the VESTA software21 unless specified otherwise. ................................................. 2

Figure 1.2. Schematic representation of the D-J n=3 RbSr2Nb3O10 crystal structure with a
ball and stick model (a) and polyhedral model (b). Pink spheres represent Rb,
Light Blue spheres represent Sr, Dark Blue and Green octahedra represent the
exterior and interior NbO6 octahedra, respectively.................................................. 4

Figure 1.3. Schematic representation of the Aurivillius phase Bi2O2(Bi2Ti3O10) crystal


structure. Red spheres = Oxygen, pink spheres = Bi, and blue octahedra = TiO6
................................................................................................................................. 5

Figure 1.4. Schematic representation of the RbCa2NaNb4O13 crystal structure. Green


spheres = Nb, Red spheres = O, Pink spheres = Rb, Blue spheres = 2/3 Ca and
1/3 Na. ..................................................................................................................... 9

Figure 1.5. Ion-exchange and intercalation behavior of layered perovskites involving


modification to interlayer gallery cations. Illustration taken from Schaak and
Mallouk (Figure 2, Page 1458).40 .......................................................................... 10

Figure 1.6. Schematic assembly processes for nanosheets through flocculation,


electrostatic sequential absorption, and finally the Langmuir-Blodgett method.
Illustration taken from Ma and Sasaki (Figure 7, Page 5091).54............................ 12

Figure 1.7. Schematic illustration of the exfoliation process using HSr2Nb3O10 and TBAOH
as the intercalating base. ........................................................................................ 13

Figure 2.1. Illustration of 27Al CQ values (in MHz) with varying degrees of symmetry and
coordination numbers. Taken from Kentgens.90 ................................................... 23

Figure 2.2. Euler angles defining the direction of B0 in the principal axis system (PAS) of
the EFG under static conditions. ............................................................................ 25

xv
Figure 2.3. Energy level diagram91 of spin I=9/2 nucleus and first- and second-order
quadrupolar contribution where ωL is the Larmor frequency. ............................... 27

Figure 2.4. a) Simulated powder pattern using only first-order quadrupolar effects
displaying all transitions for an I=9/2 nucleus with CQ = 20 MHz and η = 0 at a
magnetic field strength of 9.4 T. b) Simulated powder patterns using both first-
and second-order effects of the central transition for CQ = 20MHz with η
= 0, 0.5, and 1.0.92.................................................................................................. 28

Figure 2.5. Parameters of the Herzfeld-Berger convention as they relate to the NMR
spectrum................................................................................................................. 30

Figure 2.6. 93Nb NMR chemical shifts for NbOx polyhedral.94 ............................................... 31

Figure 2.7. 93Nb VOCS spectra for RbSr2Nb3O10 at 14.1 T with offsets of 62.5 kHz between
neighboring slices.98 The summation of the slices is given with and without
baseline correction. ................................................................................................ 32

Figure 2.8. 93Nb static and MAS spectra at 9.4 T with varying spin rates (CQ = 20 MHz, η
= 0.2, δiso = 0ppm, Ω = 500 ppm, κ = 0.5, α = β = γ = 0). Spectra were simulated
using Simpson software.110 .................................................................................... 35

Figure 2.9. WURST-N Amplitude profiles for WURST-2, WURST-4, WURST-40, and
WURST-80 pulses. ................................................................................................ 37

Figure 2.10. Schematic illustrations of WURST excitation and refocusing pulses


reproduced from Schurko et al.115 A - Single excitation pulse showing signal
attenuation during the excitation pulse. B – A refocusing pulse with 2τref = τexc
leads to removal of the second-order phase distortion. C – A refocusing pulse
with τref = τexc does not eliminate second-order phase distortions as illustrated
by the displaced echoes shown in blue and red. .................................................... 38

Figure 2.11. QCPMG Pulse Sequence.120 ................................................................................ 39

Figure 2.12. 91Zr NMR powder patterns obtained from a ZrO2 MAS rotor at 9.4 T using (a)
WURST echo and (b) WCPMG pulse sequences. Values shown on the right are
the relative signal intensities, normalized to account for the differing number
of scans. The simulation in (c) includes central transition lineshapes for the

xvi
tetragonal (dotted) and orthorhombic (dashed) phases. Illustration was
obtained from O’Dell.116 ........................................................................................ 40

Figure 2.13. BRAIN-CP pulse sequence. ................................................................................ 42

Figure 2.14. Schematic representation of the DOR sample setup requiring the spinning
about two axes simultaneously. ............................................................................. 43

Figure 2.15. Shifted-echo 3QMAS sequence with a FAM-II train of conversion pulses and
the corresponding coherence transfer pathway. ..................................................... 46

Figure 2.16. MQMAS-t2-REDOR pulse sequence. The I and S channels refer to the 1H and
93
Nb channels, respectively. An R3 recoupling sequence was used.139-140 ............ 48

Figure 4.1. XRD pattern for RbCa2Nb3O10. Le Bail fit parameters are shown in Table 4.2.
............................................................................................................................... 63

Figure 4.2. XRD pattern for RbCa1.9Sr0.1Nb3O10. Le Bail fit parameters are shown in Table
4.2. ......................................................................................................................... 64

- 64

Figure 4.3. XRD pattern for RbCa1.5Sr0.5Nb3O10. Le Bail fit parameters are shown in Table
4.2. ......................................................................................................................... 64

Figure 4.4. XRD pattern for RbCa1.4Sr0.6Nb3O10. Le Bail fit parameters are shown in Table
4.2. ......................................................................................................................... 65

Figure 4.5. XRD pattern for RbCa1.3Sr0.7Nb3O10. Le Bail fit parameters are shown in Table
4.2. ......................................................................................................................... 65

Figure 4.6. XRD pattern for RbCa1.1Sr0.9Nb3O10. Le Bail fit parameters are shown in Table
X. ........................................................................................................................... 66

Figure 4.7. XRD pattern for RbCa0.9Sr1.1Nb3O10. Le Bail fit parameters are shown in Table
4.2. ......................................................................................................................... 66

Figure 4.8. XRD pattern for RbCa0.7Sr1.3Nb3O10. Le Bail fit parameters are shown in Table
4.2. ......................................................................................................................... 67

Figure 4.9. XRD pattern for RbCa0.4Sr1.6Nb3O10. Le Bail fit parameters are shown in Table
4.2. ......................................................................................................................... 67

xvii
Figure 4.10. XRD pattern for RbSr2Nb3O10. Le Bail fit parameters are shown in Table 4.2.
............................................................................................................................... 68

Figure 4.11. XRD pattern for RbBa0.15Sr1.85Nb3O10. Le Bail fit parameters are shown in
Table 4.3. ............................................................................................................... 68

Figure 4.12. XRD pattern for RbBa0.3Sr1.7Nb3O10. Le Bail fit parameters are shown in
Table 4.3. ............................................................................................................... 69

Figure 4.13. XRD pattern for RbBa0.4Sr1.6Nb3O10. Le Bail fit parameters are shown in
Table 4.3. ............................................................................................................... 69

Figure 4.14. XRD pattern for RbBa0.5Sr1.5Nb3O10. Le Bail fit parameters are shown in
Table 4.3. ............................................................................................................... 70

Figure 4.15. XRD pattern for RbBa0.6Sr1.4Nb3O10. Le Bail fit parameters are shown in
Table 4.3. ............................................................................................................... 70

Figure 4.16. Graph of Unit Cell Volume vs. Ca content in RbCaxSr2-xNb3O10 compounds
calculated using lattice parameters from Le Bail fits. Note the gradual decrease
in unit cell volume with increasing Ca content due to the smaller ionic radius
of Ca2+ compared to Sr2+........................................................................................ 71

Figure 4.17. Graph of stacking distance in RbCaxSr2-xNb3O10 compounds vs. Ca content.


The stacking distance was obtained using the position of the (002) reflection.
Note the gradual decrease in stacking distance with increasing Ca content due
to the smaller ionic radius of Ca2+ vs. Sr2+............................................................. 71

Figure 4.18. EDS spectrum for RbCaSrNb3O10. Note the overlapping Rb and Sr signals. .. 73

Figure 4.19. SEM images of RbCa2Nb3O10 synthesized under dry microwave (a) and
molten salt (b) conditions and RbSr2Nb3O10 synthesized under dry microwave
(c) and molten salt (d) conditions. Note the increased platelet size seen in dry
microwave samples compared to molten salt samples. ......................................... 74

Figure 4.20. SEM images for attempted RbBa1.75Sr0.25Nb3O10 sample. Note the
combination of expected platelets and the presence of rod-like impurities in the
left image resulting from formation of the BaNb2O6 phase. The higher

xviii
magnification image on the right shows the rod-like nature of the impurity
particles.................................................................................................................. 75

Figure 4.23. Graphs of IR peak position vs. Ca content (Top) and Ba content (Bottom). ....... 76
93
Figure 4.24. Nb Static WURST spectra for RbSr2Nb3O10 at 9.4 T (Top) and 14.1 T
(Bottom). Blue - Experimental Spectrum; Black - Total Fit; Green - Exterior
Site ;and Red - Interior Site. ................................................................................. 78

Figure 4.25. 93Nb Static WURST spectra for RbCa0.4Sr1.6Nb3O10 at 9.4 T (Top) and 14.1 T
(Bottom). Blue - Experimental Spectrum; Black - Total Fit; Green - Exterior
Site 1 (Sr-like); Purple - Exterior Site 2 (Ca-like); and Red - Interior Site. ........ 79

Figure 4.26. 93Nb Static WURST spectra for RbCa0.5Sr1.5Nb3O10 at 9.4 T (Top) and 14.1 T
(Bottom). Blue - Experimental Spectrum; Black - Total Fit; Green - Exterior
Site 1 (Sr-like); Purple - Exterior Site 2 (Ca-like); Yellow – Impurity; and Red
- Interior Site......................................................................................................... 80

Figure 4.27. 93Nb Static WURST spectra for RbCa0.7Sr1.3Nb3O10 at 9.4 T (Top) and 14.1 T
(Bottom). Blue - Experimental Spectrum; Black - Total Fit; Green - Exterior
Site 1 (Sr-like); Purple - Exterior Site 2 (Ca-like); and Red - Interior Site. ........ 81

Figure 4.28. 93Nb Static WURST spectra for RbCa0.9Sr1.1Nb3O10 at 9.4 T (Top) and 14.1 T
(Bottom). Blue - Experimental Spectrum; Black - Total Fit; Green - Exterior
Site 1 (Sr-like); Purple - Exterior Site 2 (Ca-like); and Red - Interior Site. ........ 82

Figure 4.29. 93Nb Static WURST spectra for RbCa1.1Sr0.9Nb3O10 at 9.4 T (Top) and 14.1 T
(Bottom). Blue - Experimental Spectrum; Black - Total Fit; Green - Exterior
Site 1 (Sr-like); Purple - Exterior Site 2 (Ca-like); and Red - Interior Site. ........ 83

Figure 4.30. 93Nb Static WURST spectra for RbCa1.4Sr0.6Nb3O10 at 9.4 T (Top) and 14.1 T
(Bottom). Blue - Experimental Spectrum; Black - Total Fit; Green - Exterior
Site 1 (Sr-like); Purple - Exterior Site 2 (Ca-like); and Red - Interior Site. ........ 84

Figure 4.31. 93Nb Static WURST spectra for RbCa1.5Sr0.5Nb3O10 at 9.4 T (Top) and 14.1 T
(Bottom). Blue - Experimental Spectrum; Black - Total Fit; Green - Exterior
Site 1 (Sr-like); Purple - Exterior Site 2 (Ca-like); and Red - Interior Site. ........ 85

xix
Figure 4.32. 93Nb Static WURST spectra for RbCa1.7Sr0.3Nb3O10 at 9.4 T (Top) and 14.1 T
(Bottom). Blue - Experimental Spectrum; Black - Total Fit; Green - Exterior
Site 1 (Sr-like); Purple - Exterior Site 2 (Ca-like); and Red - Interior Site. ........ 86

Figure 4.33. 93Nb Static WURST spectra for RbCa1.9Sr0.1Nb3O10 at 9.4 T (left) and 14.1 T
(right). Blue - Experimental Spectrum; Black - Total Fit; Green - Exterior Site
1 (Sr-like); Purple - Exterior Site 2 (Ca-like); and Red - Interior Site. ............... 87
93
Figure 4.34. Nb Static WURST spectra for RbCa2Nb3O10 at 9.4 T (Top) and 14.1 T
(Bottom). Blue - Experimental Spectrum; Black - Total Fit; Purple - Exterior
Site and Red - Interior Site. .................................................................................. 88

Figure 4.35. Contour plot of 93Nb 3QMAS 2D spectrum of RbCa0.4Sr1.6Nb3O10 with double
shearing scheme. The horizontal axis corresponds to the isotropic dimension
and vertical axis to the anisotropic dimension. Note the observed signal
positions along F2 are not equal to the isotropic chemical shifts in static spectra,
as illustrated in Equation 2.42. .............................................................................. 90

Figure 4.36. Contour plot of 93Nb 3QMAS 2D spectrum of RbCa0.5Sr1.5Nb3O10 with double
shearing scheme. The horizontal axis corresponds to the isotropic dimension
and vertical axis to the anisotropic dimension. Note the observed signal
positions along F2 are not equal to the isotropic chemical shifts in static spectra,
as illustrated in Equation 2.42. .............................................................................. 90

Figure 4.37. Contour plot of 93Nb 3QMAS 2D spectrum of RbCa0.7Sr1.3Nb3O10 with double
shearing scheme. The horizontal axis corresponds to the isotropic dimension
and vertical axis to the anisotropic dimension. Note the observed signal
positions along F2 are not equal to the isotropic chemical shifts in static spectra,
as illustrated in Equation 2.42. .............................................................................. 91

Figure 4.38. Contour plot of 93Nb 3QMAS 2D spectrum of RbCa0.9Sr1.1Nb3O10 with double
shearing scheme. The horizontal axis corresponds to the isotropic dimension
and vertical axis to the anisotropic dimension. Note the observed signal
positions along F2 are not equal to the isotropic chemical shifts in static spectra,
as illustrated in Equation 2.42. .............................................................................. 91

xx
Figure 4.39. Graph of 93Nb CQ (in MHz) vs Ca content for Exterior Site 1 (blue), Exterior
Site 2 (orange), and Interior Site (grey). ................................................................ 94

Figure 4.40. Graph of 93Nb Population vs Ca content for Exterior Site 1 (blue), Exterior
Site 2 (orange), and Interior Site (grey). ................................................................ 95

Figure 4.41. Graph of 93Nb CSA span (Ω, ppm) vs Ca content for Exterior Site 1 (blue),
Exterior Site 2 (orange), and Interior Site (grey). .................................................. 95

Figure 4.42. Graph of 93Nb asymmetry parameter (η) value vs Ca content for Exterior Site
1 (blue), Exterior Site 2 (orange), and Interior Site (grey). ................................... 96

Figure 4.43. Graph of 93Nb skew (κ) value vs Ca content for Exterior Site 1 (blue), Exterior
Site 2 (orange), and Interior Site (grey). ................................................................ 96

Figure 4.44. 93Nb static WURST spectra for RbBa0.15Sr1.85Nb3O10 at 9.4 T (Top) and 14.1
T (Bottom). See Table 4.8 for detailed fit parameters. Experimental spectrum
is shown in blue and overall fit is shown in red..................................................... 99

Figure 4.45. 93Nb static WURST spectra for RbBa0.3Sr1.7Nb3O10 at 9.4 T (Top) and 14.1 T
(Bottom). See Table 4.8 for detailed fit parameters. Experimental spectrum is
shown in blue and overall fit is shown in red. ..................................................... 100

Figure 4.46. 93Nb static WURST spectra for RbBa0.4Sr1.6Nb3O10 at 9.4 T (Top) and 14.1 T
(Bottom). See Table 4.8 for detailed fit parameters. Experimental spectrum is
shown in blue and overall fit is shown in red. ..................................................... 101

Figure 4.47. 93Nb static WURST spectra for RbBa0.6Sr1.4Nb3O10 at 9.4 T (Top) and 14.1 T
(Bottom). See Table 4.8 for detailed fit parameters. Experimental spectrum is
shown in blue and overall fit is shown in red. ..................................................... 102

Figure 4.48. Graph of 93Nb CQ vs. Ba content for Exterior Site 1 (blue), Exterior Site 2
(orange), Interior Site 1 (grey), and Interior Site 2 (yellow). .............................. 105

Figure 4.49. Graph of 93Nb CQ vs. Ba content for Exterior Site 1 (blue), Exterior Site 2
(orange), Interior Site 1 (grey), and Interior Site 2 (yellow). .............................. 106

Figure 4.50. Graph of 93Nb CQ vs. Ba content for Exterior Site 1 (blue), Exterior Site 2
(orange), Interior Site 1 (grey), and Interior Site 2 (yellow). .............................. 106

xxi
Figure 4.51. Graph of 93Nb CQ vs. Ba content for Exterior Site 1 (blue), Exterior Site 2
(orange), Interior Site 1 (grey), and Interior Site 2 (yellow). .............................. 107

Figure 4.52. Graph of 93Nb CQ vs. Ba content for Exterior Site 1 (blue), Exterior Site 2
(orange), Interior Site 1 (grey), and Interior Site 2 (yellow). .............................. 107

Figure 5.1. Tetragonal crystal structure for RbCa2NaNb4O13 ................................................ 113

Figure 5.2. XRD patterns for RbCa2NaNb4O13 (above) and RbCa2NaNb2.8Ta1.2O13 (below).
Note that the pattern for the Ta compound is fit with both the four and three-
layer structures represented by blue and black ticks, respectively. ..................... 117

Figure 5.3. XRD patterns for HCa2NaNb4O13 (above) and HCa2NaNb2.8Ta1.2O13 (below).
Note that the pattern for the Ta compound is fit with both the four and three-
layer structures represented by blue and black ticks, respectively. ..................... 118

Figure 5.4. SEM images of RbCa2NaNb4O13 (top) and RbCa2NaNb2.76Ta1.24O13 (bottom)


showing platelets with lateral dimension of 1-3 μm. ........................................... 119

Figure 5.5. 93Nb Static VOCS NMR spectra with fits for HCa2NaNb4O13 at 14.1 (a) and 9.4
T (c) along with HCa2NaNb2.8Ta1.2O13 at 14.1 (b) and 9.4 T (d). The black line
corresponds to the experimental spectrum and the blue line corresponds to the
full calculated spectrum. The other colored lines correspond to individual fits
for each Nb environment. HCa2NaNb4O13 CQ values for individual fits are as
follows: purple = 22.5 MHz; red = 28.7 MHz; green = 29.5 MHz; and orange
= 52.1 MHz. For HCa2NaNb2.8Ta1.2O13 CQ values for individual fits are as
follows: purple = 22.3 MHz; red = 25.9 MHz; green = 27.1 MHz; and orange
= 51.1 MHz. ......................................................................................................... 121

Figure 5.6. 93Nb Static VOCS NMR spectra with fits for RbCa2NaNb4O13 at 9.4 (top left)
and 14.1 T (bottom left) along with RbCa2NaNb2.76Ta1.24O13 at 9.4 (top right)
and 14.1 T (bottom right). The black line corresponds to the experimental
spectrum and the blue line corresponds to the full calculated spectrum. The
other colored lines correspond to individual fits for each Nb environment. For
RbCa2NaNb4O13 CQ values, purple = 33.5 MHz, red = 35.2 MHz, green = 18.8
MHz, and orange = 53.5 MHz. For RbCa2NaNb2.8Ta1.2O13 CQ values, purple =
28 MHz, red = 33 MHz, green = 21 MHz, and orange = 51 MHz. ..................... 123

xxii
93
Figure 5.7. Nb 3QMAS 2D spectrum of RbCa2NaNb4O13 (left) and
RbCa2NaNb2.8Ta1.2O13 (right) with double shearing scheme. The horizontal
axis corresponds to the isotropic dimension and vertical axis to the anisotropic
dimension............................................................................................................. 126

Figure 5.8. 1H-93Nb BRAIN-CP spectrum with calculated fit. The black spectrum is the
experimental, blue the full calculated fit, red is the fit of site 3 (28.7 MHz), and
green is the fit of site 2 (29.5 MHz)..................................................................... 129

Figure 5.9. MQMAS-t2-REDOR spectra for HCa2NaNb4O13 without (left) and with (right)
REDOR sequence active. Numbers correspond to MQMAS sidebands for
appropriate sites. Site numbers are the same as those listed in Table 1.............. 131
23
Figure 5.9. Na MAS and MQMAS spectra for HCa2NaNb4O13 (a and b) and
HCa2NaNb2.8Ta1.2O13 (c and d). The black line corresponds to the full
experimental spectrum, the blue line is the full calculated spectrum, the red line
is the individual fit for site 1, the green line is the individual fit for site 2, and
the orange line is the individual fit for site 3. The fit parameters for each site
are listed in Table 2. ............................................................................................ 134

Figure 6.1. Illustration of stepwise microwave grafting. ....................................................... 137

Figure 6.2. (a) XRD patterns for HSr2Nb3O10 (black), C1/Sr2Nb3O10 (red), C3/Sr2Nb3O10
(green), and C6/Sr2Nb3O10 (blue) synthesized using the microwave irradiation
method. (b) XRD patterns for HCa2Nb3O10 (black), C1/Ca2Nb3O10 (red),
C3/Ca2Nb3O10 (green), and C6/Ca2Nb3O10 (blue) synthesized using a
conventional heating method. .............................................................................. 139

Figure 6.3. XRD patterns for HCa2Nb3O10 (Black), C1/Ca2Nb3O10 (red), and C3/Ca2Nb3O10
(green) synthesized using microwave grafting. Note the presence of an intense
reflection associated with the HCa2Nb3O10 phase. .............................................. 141

Figure 6.4. Long range XRD patterns for HCa2Nb3O10 (top) and HSr2Nb3O10 (bottom).
Experimental patterns are shown in blue. Le Bail fits are shown in red. Blue
ticks represent hkl reflections. The difference pattern is shown in gray. ........... 143

Figure 6.5. Long range XRD patterns for C6/Ca2Nb3O10 (top) and C6/Sr2Nb3O10 (bottom).
Experimental patterns are shown in blue. Le Bail fits are shown in red. Blue

xxiii
ticks represent hkl reflections for the C6 phase. Black ticks represent hkl
reflections for the hydrated phase. The difference pattern is shown in gray.
Listed hkl reflections refer to the C6 phase unless specified otherwise. *Refers
to the minority hydrated phase. ........................................................................... 144

Figure 6.6. (a) 1H MAS Hahn Echo NMR spectra for C6/Sr2Nb3O10 with varying echo
delay times and (b) results of peak integration as a function of the echo delay
time. ..................................................................................................................... 145

Figure 6.7. TGA data for C6/Sr2Nb3O10 (above) and C6/Ca2Nb3O10 (below) with
temperatures ranging from 100 to 800 °C. Based on theoretical values of
11.84% and 13.84% for 100% alkoxyl coverage of the strontium and calcium
samples, the alkoxyl coverage was found to be 39.1% and 45.0%, respectively.
............................................................................................................................. 146

Figure 6.8. Alkyl coverage percentages for C1-C6/CaxSr2-xNb3O10 compounds using


microwave grafting. ............................................................................................. 148

Figure 6.9. XRD patterns for H-C6/CaxSr2-xNb3O10 compounds grafted using the
microwave heating method. ................................................................................. 150

Figure 6.10. TGA graphs for C1 and C3/Sr2Nb3O10 samples. All mixed Ca/Sr samples
show the same general shape of graph with the only changes being the
magnitude of mass loss. ....................................................................................... 151

Figure 6.11. Tauc plots for H-C6/Sr2Nb3O10 (a-d, respectively) microwave grafted samples
showing substantial shift in band gap upon grafting. .......................................... 154

Figure 7.1. Image of C6/Ca2Nb3O10 in 5M2H (Sample 4) after sitting unmoved for one
week. .................................................................................................................... 164

Figure 7.2. Image of C6/Ca0.25Sr1.75Nb3O10 in 2P (Left – Sample 11), and in 5M2H (Right
– Sample 10) 30 minutes after initial sonication. Note that particles in Sample
11 are settling at a much faster rate than those in Sample 10. ............................. 165

Figure 7.3. Image of C6/Sr2Nb3O10 in 5M2H (Sample 5) shortly after microwave heating
(left) and after sitting untouched for 1 week (right). The observed Tyndall

xxiv
Effect is indicative of a colloidal suspension that does not settle out with
gravitational force alone. ..................................................................................... 166

Figure 7.4. Images for C3/Sr2Nb3O10 in 2P (Sample 8) after sitting untouched for 2 weeks
(Left) and 2 months (Right). The colloidal suspension is clearly stable after 2
months even despite approximately 50% of the solvent having evaporated. A
slight yellow tint can be observed in the sample after 2 weeks, but is ambiguous
after 2 months. ..................................................................................................... 166

Figure 7.5. Image of C6/Sr2Nb3O10 in 5M2H (Sample 15a) having sit for 5 hr after initial
sonication. Note the yellow color seen in Sample 12 from which it was
prepared is barely present, if present at all. ......................................................... 167

Figure 7.6. Image of C6/Sr2Nb3O10 in 5M2H after centrifugation (Sample 15a – Left) and
after mixing with ethanol and H2O (Sample 15b – Right). The organic layer in
Sample 15b is on top and aqueous layer on bottom. Note the substantial
cloudiness present in the aqueous layer in Sample 15b. ...................................... 168

Figure 7.7. Image of C3/Sr2Nb3O10 in 2P 2nd exfoliation SN (Sample 14c) and


C6/Sr2Nb3O10 in 5M2H 2nd exfoliation (Sample 15c). Note the yellow color in
Sample 15c returned after microwave heating .................................................... 169

Figure 7.8. Image of C6/Sr2Nb3O10 in 5M2H SN (Sample 15c) with laser beam passing
through. Observance of the Tyndall effect is evidence of a colloidal suspension.
This is the first instance in which the Tyndall Effect was observed after
centrifugation. ...................................................................................................... 169

Figure 7.9. UV-vis spectra for untreated 5M2H (Blue), 5M2H control sample (Orange),
and Nanosheet suspension (yellow)..................................................................... 170

Figure 7.10. UV-vis spectrum of C6/Sr2Nb3O10 parent compound. ...................................... 171

Figure 7.11. IR spectra for 5M2H/t-BuOH mixture untreated (Blue), 2P untreated (orange),
C6/Sr2Nb3O10 in 5M2H/t-BuOH mixture (grey), and C3/Sr2Nb3O10 in 2P
(yellow)................................................................................................................ 172
1
Figure 7.12. H NMR spectra for 5M2H untreated, 5M2H control sample, and
C6/Sr2Nb3O10 in 5M2H SN. ................................................................................ 172

xxv
Figure 7.13. XRD pattern for C6/Ca2Nb3O10 in 5M2H sediment collected from the bottom
of the sample container (Sample 4). .................................................................... 176

Figure 7.14. XRD patterns for C6/Sr2Nb3O10 in 5M2H (Sample 5) slow air-dried
supernatant (top) and the powder sediment (bottom). ......................................... 176

Figure 7.15. XRD pattern of C6/Sr2Nb3O10 in 5M2H (Sample 6) after it had sat on the
benchtop untouched for over 1 year. All solvent had evaporated leaving solid
powder on the bottom of the container. Not the C6 phase is still present. ......... 177

Figure 7.16. XRD pattern for C3/Sr2Nb3O10 in 2P SN sediment (Sample 14c). Note the
presence of both C3 and hydrated phases at 4.8 and 5.7o, respectively. .............. 177

Figure 7.17. XRD pattern for C6/Sr2Nb3O10 in 5M2H (Sample 15c) Sediment. The
stacking reflection at 5.8o is consistent with a partially hydrated phase. The
slight plateau between 3.5 and 4.7o is possibly indicative of remaining
C6/Sr2Nb3O10. ...................................................................................................... 178

Figure 7.18. XRD pattern for C6/Ca2Nb3O10 in 5M2H suspension (Sample 4) and
C6/Sr2Nb3O10 in 5M2H suspensions (Sample 5) obtained by air-drying several
drops of the suspension (before centrifugation) onto a glass plate. The
reflections at ~3.6 and ~5.8 degrees correspond to C6 grafted and hydrated
phases, respectively. Note the absence of a strong hydrated phase reflection in
Sample 5 is possibly due to differences in the suspension preparation
(sonication vs. microwave heating), differences in the procedure employed for
synthesis of the parent compound (conventional heating for Sample 4 vs.
microwave heating for Sample 5), or differing composition of the parent
compound. ........................................................................................................... 179

Figure 7.19. XRD patterns for C6/Sr2Nb3O10 in 5M2H (Sample 7) from the slow settling
phase (Top) and moderate settling phase (Bottom). ............................................ 181

Figure 7.20. XRD pattern for C3/CaSrNb3O10 in 2P (Sample 9) dried suspension. The
pattern on top is from a dried suspension prepared two weeks after the original
synthesis while the bottom pattern is from the day of the synthesis. The lack of
a grafted stacking reflection, in the dried suspension prepared after sitting two
weeks, and increased intensity of the hydrated stacking reflection suggest

xxvi
sheets in the suspension are reverting from the grafted state to the hydrated
state. ..................................................................................................................... 182

Figure 7.21. XRD pattern for C6/Sr2Nb3O10 in 5M2H centrifuge supernatant (Sample 5)
after drying. The sample was rapidly dried by placing drops of the supernatant
onto a glass plate heated to 80 oC such that solvent evaporated within seconds
until a significant amount of solid remained. The broad hump in the pattern is
due to the background signal of the glass plate (blank plate shown in blue). The
grey pattern is a simulation of (00l) stacking reflections assuming an average
number of 2 hydrated sheets stacked together. .................................................... 184

Figure 7.22. Long-range (Top) and short-range (Bottom) XRD patterns for C3/Sr2Nb3O10
in 2P (Sample 8) supernatant after sitting for 2 weeks. Only the hydrated phase
is observed with fewer higher order reflections compared to parent compound.
Note the slight hump at ~3o is due to the glass slide being place in a position
such that it blocked the X-Ray beam below 3o. ................................................... 185

Figure 7.23. Long-range XRD pattern of C6/Sr2Nb3O10 in 5M2H (Sample 13) supernatant
dried in a N2 atmosphere. The lack of many higher order reflections is
consistent with irregular restacking of sheets, though the stacking reflection
position is consistent with a hydrated phase rather than the desired grafted
phase. ................................................................................................................... 186

Figure 7.24. Long-range XRD pattern for C6/Sr2Nb3O10 after freeze-drying in a 5M2H/t-
BuOH mixture (Sample 12). The presence of a sharp stacking reflection and
higher order reflections is likely due to the frozen solution melting several times
during the freeze-drying process, essentially resulting in a slow evaporation of
solvent and thus allowed for ordered restacking of sheets. ................................. 187

Figure 7.25. SEM image of C6/Sr2Nb3O10 in 5M2H (Sample 3) on TESCAN instrument.


While the particles have the platelet shape seen in parent compounds, the
particle size is substantially smaller with 681 nm being the largest lateral
dimension seen in the sample. Plate thickness of 67 and 154 nm correspond to
~40 and ~100 sheets stacked together, respectively. ........................................... 188

xxvii
Figure 7.26. C6/Ca2Nb3O10 in 5M2H SN (Sample 4 sitting unmoved for three weeks, Left)
and C6/Sr2Nb3O10 in 5M2H SN (Sample 6, Right) SEM images. Sample was
prepared by placing several drops onto SEM sample holder and letting air dry.
Sample 6 on the right was prepared over a much longer period with ~1mL of
SN in total dried onto the Si wafer. Note the substantially increased irregularity
and decreased size of the particles compared to parent compounds. ................... 188

Figure 7.27. SEM image (top) and EDS spectrum (bottom) for C6/Sr2Nb3O10 in 5M2H
(Sample 5) supernatant after evaporated solvent, re-dissolving, centrifuging,
and final evaporation of solvent. The small and irregular shape of the particles
combined with the presence of Nb in the EDS spectrum is strong evidence for
successful exfoliation of nanosheets. Note that there is no Sr signal assigned in
the EDS spectrum due to the Si signal from the wafer overlapping. ................... 189

Figure 7.28. SEM images for (a) C3/Sr2Nb3O10 in 2P (sample 14c) SN, (b) C6/Sr2Nb3O10
in 5M2H (sample 15c) SN. Note the horizontal lines in image (a) are to due
scratches on the copper tape background created during removal of the adhesive
layer. .................................................................................................................... 190

Figure 7.30. SEM image of C6/Sr2Nb3O10 nanosheets deposited onto a Cu2O nanocube.
The small (less than 500 nm), irregularly-shaped particles are the
exfoliated/restacked nanosheets. Note that there is a wide distribution of
nanosheet agglomeration. .................................................................................... 192

Figure 7.31. SEM image of C6/Sr2Nb3O10 nanosheets deposited onto a Cu2O nanocube.
The small (less than 500 nm), irregularly-shaped particles are the
exfoliated/restacked nanosheets. Note that the almost transparent quality of
nanosheets on the cube surface is consistent with the electron beam passing
through the thin sheet. ......................................................................................... 193

Figure A.1. List of indexed hkl reflections obtained from XRD pattern of RbSr2Nb3O10
microwave sample. .............................................................................................. 219

Figure A.2. List of indexed hkl reflections obtained from XRD pattern of RbSr2Nb3O10
molten salt sample ............................................................................................... 220

xxviii
Figure A.3. List of indexed hkl reflections obtained from XRD pattern of
RbCa0.3Sr1.7Nb3O10............................................................................................... 221

Figure A.4. List of indexed hkl reflections obtained from XRD pattern of
RbCa0.4Sr1.6Nb3O10............................................................................................... 221

Figure A.5. List of indexed hkl reflections obtained from XRD pattern of
RbCa0.5Sr1.5Nb3O10............................................................................................... 222

Figure A.7. List of indexed hkl reflections obtained from XRD pattern of
RbCa0.7Sr1.3Nb3O10............................................................................................... 223

Figure A.8. List of indexed hkl reflections obtained from XRD pattern of
RbCa1.1Sr0.9Nb3O10............................................................................................... 224

Figure A.9. List of indexed hkl reflections obtained from XRD pattern of
RbCa1.4Sr0.6Nb3O10............................................................................................... 226

Figure A.10. List of indexed hkl reflections obtained from XRD pattern of
RbCa1.9Sr0.1Nb3O10............................................................................................... 227

Figure A.11. List of indexed hkl reflections obtained from XRD pattern of RbCa2Nb3O10 .. 228

Figure A.12. List of indexed hkl reflections obtained from XRD pattern of
RbBa0.15Sr1.85Nb3O10 ............................................................................................ 229

Figure A.13. List of indexed hkl reflections obtained from XRD pattern of
RbBa0.3Sr1.7Nb3O10............................................................................................... 230

Figure A.14. List of indexed hkl reflections obtained from XRD pattern of
RbBa0.6Sr1.4Nb3O10............................................................................................... 231

Figure B.1. TGA data for C1/Sr2Nb3O10 with temperatures ranging from 150 to 800 °C.
See Table B.1 for mass loss % and coverage %. ................................................. 233

Figure B.2. TGA data for C3/Sr2Nb3O10 with temperatures ranging from 150 to 800 °C.
See Table B.1 for mass loss % and coverage %. ................................................. 233

Figure B.3. TGA data for C6/Sr2Nb3O10 with temperatures ranging from 150 to 800 °C.
See Table B.1 for mass loss % and coverage %. ................................................. 234

xxix
Figure B.4. TGA data for C1/Ca0.4Sr1.6Nb3O10 with temperatures ranging from 150 to 800
°C. See Table B.1 for mass loss % and coverage %. ........................................... 234

Figure B.5. TGA data for C3/Ca0.4Sr1.6Nb3O10 with temperatures ranging from 150 to 800
°C. See Table B.1 for mass loss % and coverage %. ........................................... 235

Figure B.6. TGA data for C6/Ca0.4Sr1.6Nb3O10 with temperatures ranging from 150 to 800
°C. See Table B.1 for mass loss % and coverage %. ........................................... 235

Figure B.7. TGA data for C1/Ca1Sr1Nb3O10 with temperatures ranging from 150 to 800 °C.
See Table B.1 for mass loss % and coverage %. ................................................. 236

Figure B.8. TGA data for C3/Ca1Sr1Nb3O10 with temperatures ranging from 150 to 800 °C.
See Table B.1 for mass loss % and coverage %. ................................................. 236

Figure B.9. TGA data for C6/Ca1Sr1Nb3O10 with temperatures ranging from 150 to 800 °C.
See Table B.1 for mass loss % and coverage %. ................................................. 237

Figure B.10. TGA data for C1/Ca1.5Sr0.5Nb3O10 with temperatures ranging from 150 to 800
°C. See Table B.1 for mass loss % and coverage %. ........................................... 237

Figure B.11. TGA data for C3/Ca1.5Sr0.5Nb3O10 with temperatures ranging from 150 to 800
°C. See Table B.1 for mass loss % and coverage %. ........................................... 238

Figure B.12. TGA data for C6/Ca1.5Sr0.5Nb3O10 with temperatures ranging from 150 to 800
°C. See Table B.1 for mass loss % and coverage %. ........................................... 238

Figure B.13. TGA data for C1/Ca1.8Sr0.2Nb3O10 with temperatures ranging from 150 to 800
°C. See Table B.1 for mass loss % and coverage %. ........................................... 239

Figure B.14. TGA data for C3/Ca1.8Sr0.2Nb3O10 with temperatures ranging from 150 to 800
°C. See Table B.1 for mass loss % and coverage %. ........................................... 239

Figure B.15. TGA data for C6/Ca1.8Sr0.2Nb3O10 with temperatures ranging from 150 to 800
°C. See Table B.1 for mass loss % and coverage %. ........................................... 240

Figure B.16. TGA data for C6/Ca2Nb3O10 with temperatures ranging from 150 to 800 °C.
See Table B.1 for mass loss % and coverage %. ................................................. 240

xxx
List of Abbreviations

TBAOH: Tetrabutylammonium hydroxide

5M2H: 5-methyl-2-hexanone

2P: 2-pentanone

C1: methanol grafted

C3: n-propanol grafted

C6: n-hexanol grafted compound

NMR: Nuclear Magnetic Resonance

SSNMR: Solid-State Nuclear Magnetic Resonance

MQ: Multiple Quantum

PAS: Principle Axis System

MAS: Magic Angle Spinning

WURST: Wideline Uniform Rate Smooth Truncation

BRAIN: Broadband Adiabatic Inversion

CP: Cross-Polarization

REDOR: Rotational Echo Double Resonance

VOCS: Variable Offset Cumulative Spectra

FAM: Fast Amplitude Modulated

SN: Supernatant

RF: Radiofrequency

IR: Infrared

xxxi
TGA: Thermal Gravimetric Analysis

XRD: X-Ray Diffraction

EFG: Electric Field Gradient

CSA: Chemical Shift Anisotropy

D-J: Dion-Jacobson

R-P: Ruddlesden-Popper

SEM: Scanning Electron Microscopy

EDS: Energy Dispersive Spectroscopy

CTC: Charge-Transfer Complex

MSS: Molten Salt Synthesis

CPMG: Carr-Purcell-Meiboom-Gill

xxxii
Chapter 1: Background of Layered Perovskites

1.1 Introduction

The perovskite structure can be synthesized in such variety as to produce a wide

array of phases with vastly differing functions such as capacitance,1-3 piezoelectricity,4-6

insulation,7-9 metallic conduction,10-12 catalysis,13-15 and superconduction16-17 to name a

few. Furthermore, in just the past five years, hybrid organic-inorganic perovskites have

rapidly gained stardom to become the most studied semiconductors for use in photovoltaic

cells.18-19 In 1839, CaTiO3 was discovered in the Ural Mountains and subsequently named

Perovskite after Russian mineralogist Lev A. Perovski. 177 years later the term perovskite

is used to describe the large class of materials consisting of the same ABX3 structure.

Goldschmidt et al.20 performed pioneering work in the 1920’s on the perovskite structure

that laid the foundation for further exploration of the perovskite family which has bloomed

into arguably the most diverse class of ceramic materials. Even after over 100 years of

active research into perovskite compounds, there is no apparent slowdown in active

research as illustrated by 522 ACS published articles, in 2016 alone, containing the word

perovskite in the title. Such is the beauty of scientific research that the lone seedling of

discovery gives rise to a great tree of knowledge and understanding with branches reaching

out far beyond what possibly could have been envisioned when planted.

As previously mentioned, the chemical formula for perovskites is given as ABX3

with A being a Group I or II cation, B is a metal, and X is typically oxygen, though other

ions such as F- or Cl- are possible. SrTiO3 represents an idealized cubic structure with Sr2+

and O2- forming a closed packed cubic lattice with Ti4+ occupying the octahedral holes and

1
resulting in a three-dimensional structure of corner sharing TiO6 octahedra with Sr2+ ions

occupying the cavities between octahedra as shown in Figure 1.1. Both A and B cations

can be varied to a great degree resulting in a myriad of compounds with perovskite

structures.

Figure 1.1. Schematic representation of the ideal perovskite structure based on the cubic SrTiO 3
crystal structure. Left- Ball and stick model of unit cell where blue spheres = Ti, red spheres = O, and
green sphere = Sr. Right- polyhedral model showing TiO6 octahedra in blue and Sr occupying
interstitial 12-coordinate sites in green. All crystallographic representations have been generated
using the VESTA software21 unless specified otherwise.

The fundamental lattice parameter, a, in an ideal cubic perovskite can be calculated

using ionic radii rA, rB, and rO as given by the following equation:

Equation 1.1:𝑎 = √2(𝑟𝐴 + 𝑟𝑂 ) = 2(𝑟𝐵 + 𝑟𝑂 )

It is possible to estimate the degree of distortion by taking the ratio of the two above

expressions to give Goldschmidt’s tolerance factor, t, calculated using Eq. 1.2. While this

tolerance factor assumes purely ionic bonding, it serves as a good approximation for

compounds with high degrees of ionic bonding.

2
𝑟𝐵 +𝑟𝑂
Equation 1.2: 𝑡 =
√2(𝑟𝐴 +𝑟𝑂 )

The cubic structure is maintained for values of 0.89 < t < 1 with BO6 octahedra

tilting in order to maintain connectivity and minimize void space.22 For lower values of t,

the resulting crystal structure is stable in a lower symmetry environment. At the other end

of the spectrum, if t is greater than 1, hexagonal variants of the perovskite structure are

stable resulting in face sharing BO6 octahedra. It should be noted again that t serves as an

approximation and there are exceptions to these rules. Furthermore, distortions of the

perovskite structure can be caused by varying chemical composition and Jahn-Teller

distortions.

Among the class of perovskite materials, layered perovskites have arguably been

the most heavily studied in recent years due to a myriad of properties. There are three main

classes of layered perovskites, each consisting of the common motif of two-dimensional

anionic perovskite slabs stacked together with interleaving cations or cationic structural

units between each slab. These three phases of layered perovskite are the Aurivillius,

Ruddlesden-Popper (R-P), and Dion-Jacobson (D-J) phases. The D-J series of layered

perovskites, A’[An-1BnO3n+1] where n corresponds to the number of B-site octahedra, is

typified by the n = 3 triple layer RbSr2Nb3O10 (Figure 1.2) with one Rb interlayer cation

per Sr2Nb3O10 formula unit. A’ corresponds to the interlayer cation (in this case a group I

cation), A is the interstitial site (typically a group II element, although other +2 cations can

be suitable if they have an appropriate ionic radius size match), and B is the metal center

for the BO6 octahedra (typically a +5 metal for D-Js). R-P phases, A’2[An-1BnO3n+1], such

3
as K2La2Ti3O10 have two interlayer cations per formula unit and possess twice the

interlayer charge density of the D-J phases. The B site in the R-P phase is typically a +4

metal such as Ti4+. Aurivillius phases (see Figure 1.3), including Bi2W2O9, are

intergrowths of perovskite and bismuth oxide with a covalent network of Bi2O2+ between

the two-dimensional perovskite slabs.

a) b)

Figure 1.2. Schematic representation of the D-J n=3 RbSr2Nb3O10 crystal structure with a ball and
stick model (a) and polyhedral model (b). Pink spheres represent Rb, Light Blue spheres represent Sr,
Dark Blue and Green octahedra represent the exterior and interior NbO 6 octahedra, respectively.

4
Figure 1.3. Schematic representation of the Aurivillius phase Bi2O2(Bi2Ti3O10) crystal structure. Red
spheres = Oxygen, pink spheres = Bi, and blue octahedra = TiO6

Layered niobates have been of particular interest to researchers, due to the

versatility of Nb occupying the B-site. Different oxidation states (the most common and

stable being Nb(V)), along with high charges and small ionic radii creating a strong oxygen

affinity for Nb, has led to an extensive number of different niobate structures, with

elemental composition leading to a wide variety of interesting properties. The layered

niobates studied in this work belong entirely to the D-J class of layered perovskites.

Perhaps the most useful property of layered niobates is their inherent ability to undergo

soft-chemical structural modifications, specifically within the interlayer gallery separating

the perovskite blocks.

5
1.2 RbCaxSr2-xNb3O10 and RbCa2NaNb4-xTaxO13 Crystal Structures

A significant portion of this dissertation involves the n=3 RbCaxSr2-xNb3O10 and

n=4 RbCa2NaNb4-xTaxO13 versions of D-J perovskites, and thus it makes logical sense to

devote some time to an in-depth discussion of the synthesis and crystal structures for these

compounds derived from previous work before diving into the research conducted here. It

should be noted that while a great amount of study has been done into RbCa2Nb3O10,

RbSr2Nb3O10, and RbCa2NaNb4O13 structures, little information is available on the mixed

cation structures outside of the work presented herein.

RbCa2Nb3O10 was originally synthesized in 1981 by M. Dion23 and was initially

thought to form in the tetragonal space group P4212 with lattice constants a = 7.725(2) Å

and c = 14.909(5) Å. The original synthesis called for the heating of the carbonate and

oxide precursors Rb2CO3, CaCO3, and Nb2O5 in a 1.25:2:3 molar ratio for Rb:Ca:Nb (the

slight excess of Rb2CO3 is due to its volatility) at 1300 oC in a furnace for 24 hours, which

is the standard method of perovskite synthesis. Liang et al.24 later refined the crystal

structure from X-ray powder diffraction data using Rietveld refinement and found it to

exist in the higher symmetry tetragonal P4/mmm space group with a=3.85865(6) Å and

c=14.9108(3) Å. An alternative method involving substantially lower temperatures known

as molten salt synthesis (MSS)25-28 (discussed further in Chapter 3) has been explored for

many layered perovskites with success. Geselbracht et al.29 applied this method to

RbCa2Nb3O10 and found it could be readily synthesized in a RbCl flux at 800 oC in only 1

hour while maintaining the same tetragonal symmetry observed in conventional syntheses.

6
The NbO6 octahedra have substantial distortions resulting in a tilting effect due to the ionic

radii of Ca2+ not completely filling the void space at the interstitial site.

Thangadurai et al.30 reported the crystal structure for RbSr2Nb3O10 in the tetragonal

space group P4/mmm with a = 3.8944(2) Å and c = 15.2710(8) Å. Sr2+ has a larger ionic

radius than Ca2+ thus resulting in less octahedral tilting and an expansion of the crystal

lattice that is most dramatically seen in the difference in the c lattice parameters in

RbCa2Nb3O10 and RbSr2Nb3O10. As with its Ca counterpart, RbSr2Nb3O10 can be

synthesized using either conventional high temperature synthesis or lower temperature

MSS. Geselbracht et al.31 synthesized a series of RbCaxSr2-xNb3O10 compounds using the

molten salt method and found the lattice parameters in the tetragonal unit cell progressed

in a linear fashion (see Table 1.1).

Table 1.1. Tetragonal unit cell parameters of RbCaxSr2-xNb3O10 phases collected by Geselbracht et al.31
X Value a (Å) c (Å) V (Å3)
2 3.8551(8) 14.871(5) 221.01(9)
1.8 3.8598(9) 14.906(5) 222.1(1)
1.6 3.8632(8) 14.952(6) 223.2(1)
1.4 3.8664(9) 14.995(5) 224.2(1)
1.2 3.8704(9) 15.040(5) 225.3(1)
1.0 3.874(1) 15.078(5) 226.3(1)
0.8 3.8771(8) 15.116(4) 227.2(1)
0.6 3.8801(8) 15.149(4) 228.1(1)
0.4 3.8834(9) 15.178(5) 228.9(1)
0.2 3.8864(9) 15.208(5) 229.7(1)
0 3.8897(9) 15.233(5) 230.4(1)

The Nb-O bond lengths for RbSr2Nb3O10 as determined by Wang et al.32are shown

in Table 1.2, with site assignments illustrated in Figure 1.2a. The axial terminal Nb2-O4

bond length shown in Table 1.2 is substantially shortened at 1.7525 Å, to the point where

this bond can be considered essentially a Nb-O double bond with Nb in a pseudo five

coordinate environment. Complementing this contraction of the terminal Nb2-O4 bond,

7
the Nb2-O2 bond bridging the exterior and interior octahedra is substantially elongated to

2.4539 Å from the ideal bond length of 2.06 Å calculated based on Shannon-Prewitt Crystal

Radii.33 Elongation of the axial bridging B-O bond along with shortening of the axial

terminal B-O bond is characteristic of all alkali layered D-J perovskites.

Table 1.2 Nb-O bond lengths in RbSr2Nb3O10.


Bond Length (Angstroms)
Nb1(Interior)-O2(Axial) 1.95202
Nb1(Interior)-O1(Equatorial) 2.00852
Nb2(Exterior)-O2(Axial) 2.45386
Nb2(Exterior)-O3(Equatorial) 1.98825
Nb2(Exterior)-O4(Axial) 1.75245

KCa2NaNb4O13 was first synthesized by Jacobson et al.34 through the incorporation

of an additional perovskite layer of NbO6 octahedra onto the parent compound

KCa2Nb3O10. RbCa2NaNb4O13 was synthesized by mixing equimolar amounts of

RbCa2Nb3O10 and NaNbO3 followed by heating at 1200 oC for 3 hours. Sato et al.35 solved

the crystal structure for this compound, shown in Figure 1.4, using Rietveld analysis of

powder X-ray diffraction data and found it exists in the tetragonal P4/mmm space group

with a=3.8702(1) Å and c=18.8935(6) Å where Na+ and Ca2+ are uniformly distributed

among the interstitial sites. A thorough literature search has found no mention of mixed

Nb/Ta RbCa2NaNb4-xTaxO13 compounds such as those studied in this work. As will be

discussed at length in later chapters, the crystal structure proposed by Sato et al.35 does not

correspond to a stable structure as evidenced by bond valence calculations nor does it agree

with results from NMR experiments.

8
Figure 1.4. Schematic representation of the RbCa2NaNb4O13 crystal structure. Green spheres = Nb,
Red spheres = O, Pink spheres = Rb, Blue spheres = 2/3 Ca and 1/3 Na.

1.3 Soft-chemical Exchange Reactions in Layered Perovskites

A particularly desirable property of layered perovskites is the ability to undergo ion

exchange to prepare a wide variety of novel materials using low-temperature reaction

conditions. The most common form of ion exchange is the acid exchange reaction where

the layered perovskite is treated with a strong acid thus exchanging the interlayer cations

with acidic protons. Acid exchanging has led to many innovations36-39 regarding the

catalytic ability of layered perovskites due to the generation of a solid acid after successful

exchange. Of course, for these active acid sites to be any of use, the reactant must first

reach the active site and in the layered form; unless the reactant can enter the interlayer

gallery these sites are useless. Therefore, it is highly desirable to exfoliate layered niobates

9
into two-dimensional nanosheets. The most common route towards exfoliation into

nanosheets is through intercalation of bulky organic bases such as tetrabutylammonium

hydroxide (TBAOH) into the protonated interlayer gallery. Schaak and Mallouk40 have

created an illustrated and informative summary of ion-exchange and intercalation behavior

of layered perovskites which is shown in Figure 1.5.

Figure 1.5. Ion-exchange and intercalation behavior of layered perovskites involving modification to
interlayer gallery cations. Illustration taken from Schaak and Mallouk (Figure 2, Page 1458).40

10
1.4 Niobate Nanosheets

Along with the obvious benefit of increased surface area and access to catalytically

active acid sites, nanosheets can also serve as building blocks for novel materials through

controlled reassembly of sheets.41-42 An excellent example of the potential for 2D

nanosheets as building blocks comes from Sasaki et al.43 where it was found that through

a layer-by-layer assembly approach using Langmuir-Blodgett depositions, superlattices of

(LaNb2O7/Ca2Nb3O10)5 and (Sr2Nb3O10/Ca2Nb3O10)5 possessed properties such as

increased dielectric constants and piezoelectric responses not seen in the bulk materials. A

visual guide to the Langmuir-Blodgett assembly of nanosheets is shown in Figure 1.6.

Furthermore, nanosheets have shown incredible promise in areas of photovoltaic cells,44-46

nanodielectrics,47-48 photocatalysis,49 solid acid catalysis,50-52 and flexible memory

devices53 among others. Unfortunately, it is especially difficult to characterize nanosheets

(an essential step in the development of new materials and new applications for old

materials) and it is thus of great importance to successfully characterize the parent

compounds to elucidate the structure and behavior of the resulting nanosheets.

11
Figure 1.6. Schematic assembly processes for nanosheets through flocculation, electrostatic sequential
absorption, and finally the Langmuir-Blodgett method. Illustration taken from Ma and Sasaki (Figure
7, Page 5091).54

The typical exfoliation method requires further exchanges or intercalations to be

performed on the protonated niobates to expand the interlayer gallery to a point that the

perovskite slabs break free from each other and single nanosheets are produced. These

acid-exchanged perovskites can intercalate a variety of organic bases such as alkylamines.

Intercalation with a bulky base such as tetrabutyl ammonium hydroxide (TBAOH) expands

the interlayer gallery to the point of exfoliation into single sheets. While this method of

exfoliation has been tried and true, it is decidedly lacking in many respects. The reaction

time required (typically 6 days minimum for effective exfoliation) severely inhibits the rate

at which novel compounds can be studied. Furthermore, this method is limited to using

aqueous solvents and the removal of the TBAOH cations with acid to regenerate the acid

12
sites without restacking of the sheets is an incredibly sensitive procedure that requires the

near constant presence of the investigator over the course of several hours to ensure acid

addition does not go above a certain rate. As trivial as complaints of procedural tediousness

may appear, this labor-intensive approach acts as an inhibitor to research pursuits. A

schematic illustration of the exfoliation process using TBAOH is shown in Figure 1.7.

Figure 1.7. Schematic illustration of the exfoliation process using HSr2Nb3O10 and TBAOH as the
intercalating base.

An alternative to the traditional exfoliation procedure will be discussed in great

depth in this dissertation. This new method of exfoliation involves first the grafting of long

chain alcohols into the interlayer gallery followed by treatment with an appropriate organic

solvent to break sheets apart. As will be shown, this new exfoliation method diversifies

potential reaction conditions as well as substantially shortening overall synthetic time.

While X-ray diffraction (XRD) has been used extensively in the study of layered

perovskite structure, it provides only information about long range order and is not well

suited for the study of small structural changes. Nuclear Magnetic Resonance (NMR),

however, is sensitive to the local environments surrounding each nucleus of interest and is

13
an excellent tool for the study of slight structural modifications that do not exhibit long-

range order. Signals arising from unique nuclear environments within the crystal structure

further make NMR an appropriate technique for determining the relationship between

structure and composition. In this regard, this dissertation will detail the extensive solid
93
state Nb NMR study of mixed alkaline cation layered perovskites Rb[AxSr2-xNb3O10]

where A is either Ca or Ba.

1.5 Research Presented Herein

The research described in this dissertation spans nearly six years and can be broken

into three heavily related research projects:

4. Exploration of local structural changes in mixed alkali layered perovskites.

5. Determination of B-site preference in mixed Nb/Ta systems using solid-state NMR

6. Microwave assisted grafting and exfoliation of layered perovskites using long chain

alcohols.

1.5.1 Exploration of local structural changes in mixed alkali layered perovskites.

A series of three-layered D-J RbSrxCa2-xNb3O10 (0 < x < 2) and RbBaxSr2-xNb3O10

(0 < x < 0.8) compounds were synthesized using microwave and molten salt synthetic

routes, respectively, and analyzed using powder x-ray diffraction (XRD), scanning electron

microscopy (SEM), energy dispersive spectroscopy (EDS), attenuated total reflectance-

infrared spectroscopy (ATR-IR), and wideline uniform rate smooth truncation quadrupolar

Carr-Purcell-Meiboom-Gill (WURST-QCPMG) NMR. XRD patterns yielded lattice

parameters consistent with the expansion of the crystal lattice with increasing Sr or Ba

14
content due to the size of the Sr and Ba cations. ATR-IR spectra indicated a linear

progression in the band shift from 768.9 to 738.2 cm-1 going from the pure calcium

compound to the pure strontium compound and from 738.2 to 691.5 cm-1 going from the

pure strontium compound to the x = 0.6 barium containing compound. NMR studies were

used to obtain electric field gradient (EFG) and chemical shift anisotropy (CSA)

information for the 93Nb sites in the compounds to correlate change in the local structure

with change in the A-site cations. The EFG value was found to decrease with increasing Sr

content in the Ca/Sr series and increasing Ba content in the Sr/Ba series indicating a

reduction in NbO6 octahedral distortions, specifically octahedral tilt.

1.5.2 Determination of B-site preference in mixed Nb/Ta systems using solid-state NMR

D-J phase layered perovskites can readily undergo proton exchange under mild

conditions leading to the formation of a solid acid capable of intercalating various organic

molecules.40 The B-site atom in this class of compounds is often Ti4+, Nb5+, or Ta5+. It

has been shown that changing the B-site atom can drastically alter the intercalation

behavior and catalytic ability of these materials.31, 55-56 Geselbracht et al.31 studied the

acidity of HCa2NbxTa3-xO10 compounds through intercalation and found that the pyridine

intercalation reaction did not work for x = 2.5 and 3.0 compounds suggesting a significant

decrease in solid acid strength. For x = 0.5-2.0, however, the level of pyridine intercalation

was greater than in the x=0 compound suggesting the ordering of Ta5+ within the structure

has an effect as to whether the solid acid strength is decreased or increased. In work on

the water splitting ability of SrTa2O7 and SrNb2O7, Kudo et al.57 found the water splitting

15
activity and band gaps of the SrTa2O7 and SrNb2O7 compounds to be vastly different (4.6

eV vs. 3.9 eV, respectively). In their work, no research was done on mixed Nb5+/Ta5+

systems.

While Nb5+ and Ta5+ possess the same charge and approximately equal ionic radii58

(0.64 Ǻ), Nb is more electronegative than Ta (1.6 vs. 1.5, respectively)59 leading to Nb-O

bonds, which are more covalent in character than Ta-O bonds and changes in the energy

states of Ta5+ and Nb5+ compounds.57, 60 Studies on site ordering in mixed Nb5+/Ti4+ layered

perovskites61-62 have shown site preference based on the different charges and ionic radii

(0.64 vs. 0.605 Ǻ for Nb5+ and Ti4+, respectively) of Nb5+ and Ti4+. As subsequent chapters

will show, in mixed Nb5+/Ta5+ systems, electronegativity is a strong driver of site ordering.

As illustrated in Figure 1.4, there are two distinct crystallographic B-sites in the

four-layer RbCa2NaNb4-xTaxO13 system, two interfacial octahedra, Nb(2), and two interior

octahedra, Nb(1). X-ray powder diffraction is unable to give quantitative results as to the

populations of Nb5+ and Ta5+ at each crystallographic site due to the disordered nature of

these compounds. NMR, however, allows for the study of the local structure and is ideal

for systems of this sort. Cation site-ordering in these systems will be discussed in detail in

Chapter 5.

1.5.3 Microwave assisted grafting and exfoliation of layered perovskites using long chain
alcohols.
Effective catalytic behavior in these compounds requires access to the protonated

environments typically achieved through exfoliation into single nanosheets.63 Sugahara et

al.64-69 has pioneered grafting techniques for attachment of various organic groups into the

16
interlayer of layered perovskites. In particular, Sugahara has demonstrated a grafting

technique for protonated D-J perovskites to produce derivatives with n-alcohols as large as

n-octadecanol. Current grafting methods of layered perovskites involve high pressure

heating at 80 to 150 oC for days to weeks, depending on the alcohol involved. The long

reaction time involved inhibits the ability to study grafted samples in a timely fashion. In

addition to providing more detailed background, Chapters 6 and 7 in this dissertation will

present new microwave-assisted grafting and exfoliation techniques, along with detailed

comparisons to conventional methods, designed to expand and improve current

methodologies.

17
Chapter 2: NMR Background

2.1 Solid-State NMR

In 1946, NMR was first reported for use in bulk materials by Bloch, Hansen, and

Packard70 and by Purcell, Torrey, and Pound.71 In just six years, this discovery was

recognized with the Nobel Prize in Physics being awarded to Block and Purcell. In the

subsequent seven decades, NMR has evolved into arguably the most important

spectroscopic characterization technique especially after the implementation of

superconducting magnetics which have significantly reduced the problems of magnetic

field inhomogeneity along with the development of Fourier Transform NMR (FT-NMR)

by Richard Ernst72 which resulted in yet another NMR related Nobel Prize in 1991 for

Chemistry. Furthermore, NMR was adapted for use in biological imaging resulting in the

robust diagnostic method of Magnetic Resonance Imaging (MRI) which itself resulted in

the 2003 Nobel Prize in Physiology or Medicine being awarded to Lauterbur73 and

Mansfield.74

Liquid samples generally yield sharp NMR signals due to rapid isotropic molecular

tumbling averaging the anisotropic NMR interactions to zero. In solid samples, however,

the effects of anisotropic interactions are not averaged out thus resulting in substantial

broadening of the observed signals. While this signal broadening does complicate

extraction of individual NMR interactions, solid-state NMR (SSNMR) is still an incredibly

useful tool for studying local structure (as opposed to the observation of long range periodic

structure in X-Ray diffraction). The development of numerous techniques to overcome the

issues of line broadening have allowed for SSNMR to be utilized in the studies of a variety

18
of materials such as crystalline,75 non-crystalline,76 and porous materials.77 SSNMR has

been used in the examination of complex organic and inorganic systems,78 biomolecular

systems,79 and polymeric systems.80

When NMR active nuclei are placed in a magnetic field, B0, their energy levels split

and shift due to the Zeeman effect. Subsequent application of a radiofrequency (rf) pulse

induces transitions between energy levels. The Zeeman interaction is treated as the

dominant interaction and is described by the Hamiltonian:

̂0 = −𝜇̂ 𝐵0
Equation 2.1: 𝐻

where 𝜇̂ is the nuclear magnetic moment operator. Since B0 is generally applied along the

z-axis, the Hamiltonian can be rewritten in terms of the spin operator 𝐼̂𝑍 :

̂0 = −𝛾ħ𝐼̂𝑍 𝐵0 = −ħ𝐼̂𝑍 𝜔𝐿
Equation 2.2: 𝐻

where ωL is the Larmor frequency (i.e. the rate of precession of the magnetic moment

around the external magnetic field), ωL = γB0, in units of Hz, ħ is the reduced Planck

constant, and γ is the gyromagnetic ratio.

̂ of the spin system is a


According to perturbation theory, the total Hamiltonian 𝐻

summation of the Zeeman interaction along with chemical shift, dipolar coupling, and

quadrupolar coupling and can be described as:

̂=𝐻
Equation 2.3: 𝐻 ̂0 + 𝐻
̂𝐶𝑆 + 𝐻
̂𝑑 + 𝐻
̂𝑄

19
Depending on the magnitude of the various interactions, the perturbation terms may require

higher-order corrections. As will be shown in later sections, the quadrupolar Hamiltonian

̂𝑄(2) contribution for the 93Nb spin systems


requires the consideration of the second order 𝐻

studied in this work.

The magnetic moments of nuclear spins can interact through space resulting in the

dipolar coupling interaction. Dipolar coupling is a major contributor to line-broadening in

the solid-state. The Hamiltonian for dipolar coupling between two spins I and S (given in

angular frequency units, rad/s) is given by the following equation:

𝜇 𝛾𝐼 𝛾𝑆 ħ
̂𝑑𝑑
Equation 2.4: 𝐻 ℎ𝑒𝑡𝑒𝑟𝑜
= − (4𝜋0 ) 𝑟3
𝐼̂𝑍 𝑆̂𝑍 (3𝑐𝑜𝑠 2 𝜃 − 1)

where r3 is the internuclear distance between spins I and S.

As will be discussed further in later sections, under Magic Angle Spinning (MAS)

conditions the dipolar interaction is reduced to zero. However, many solid-state

techniques81-85 have been developed which allow for dipolar coupling to be reintroduced

under MAS by applying rf pulses to modify subsequent Hamiltonians thus allowing for the

extraction of information regarding internuclear distances.

2.2 Quadrupolar Nuclei

While most are familiar with NMR only using spin ½ nuclei, 78% of NMR active

nuclei have integer (6%) or half-integer (72%) quadrupolar spins. Quadrupolar nuclei

possess an electric quadrupole moment, eQ, that interacts with the electric field gradient

(EFG). The EFG is created by the distribution of charges surrounding the nucleus being

20
studied and is directly proportional to the asymmetry of the charges, with a more

asymmetric charge distribution creating a larger EFG. This interaction leads to broadening

of NMR powder patterns and complicates analysis such that more advanced experiments

are needed, although it also provides useful information relating to the local environments

of nuclei.

The quantitative description of the EFG is presented as a 3x3 tensor, V, shown in

Equation 2.5, where the components Vij relate to the three-dimensional distribution of

charges around the nucleus:

𝑉𝑋𝑋 𝑉𝑋𝑌 𝑉𝑋𝑍


Equation 2.5: 𝑉 = [𝑉𝑋𝑌 𝑉𝑌𝑌 𝑉𝑌𝑍 ]
𝑉𝑋𝑍 𝑉𝑌𝑍 𝑉𝑍𝑍

Components Vij can be approximated using a rudimentary point charge model using

Equation 2.6, where qe is nearby charge, r is the distance from the nucleus to the charge,

and x, y, and z are the bond distance components along the various axes:

3𝑥𝑖2 −𝑟𝑖2 3𝑥𝑖 𝑦𝑖 3𝑥𝑖 𝑧𝑖


∑𝑖 𝑞𝑖 ( ) ∑𝑖 𝑞𝑖 ( ) ∑𝑖 𝑞𝑖 ( )
𝑟𝑖5 𝑟𝑖3 𝑟𝑖3
3𝑥𝑖 𝑦𝑖 3𝑦𝑖2 −𝑟𝑖2 3𝑦𝑖 𝑧𝑖
Equation 2.6: 𝑉 = 𝑒 ∑𝑖 𝑞𝑖 ( 𝑟𝑖3
) ∑𝑖 𝑞𝑖 (
𝑟𝑖5
) ∑𝑖 𝑞𝑖 (
𝑟𝑖3
)
3𝑥𝑖 𝑧𝑖 3𝑦𝑖 𝑧𝑖 3𝑧𝑖2 −𝑟𝑖2
∑𝑖 𝑞𝑖 ( ) ∑𝑖 𝑞𝑖 ( ) ∑𝑖 𝑞𝑖 ( )
[ 𝑟𝑖3 𝑟𝑖3 𝑟𝑖5 ]

The EFG tensor is described in its own principal axis system (PAS) through

diagonalization:

𝑉𝑋𝑋 0 0
Equation 2.7: 𝑉 = [ 0 𝑉𝑌𝑌 0 ]
0 0 𝑉𝑍𝑍

21
The quadrupolar Hamiltonian for spin I in applied magnetic field B 0 is given by:86-
88

𝑒𝑄
Equation 2.8: Ĥ𝑄 = 𝑰∙ 𝑽∙𝑰
6𝐼(𝐼−1)ħ

The terms are typically defined such that |Vxx| < |Vyy| < |Vzz| and Vxx + Vyy + Vzz =

0.89 The quadrupolar coupling constant (CQ) is directly related to the EFG as shown

in Equation 2.9. CQ refers to the interaction of the quadrupole moment, eQ, and the

electric field gradient, eVzz. Another EFG related term involved in quadrupolar

nuclei is the asymmetry parameter, η, shown in Equation 2.10 with values ranging

from 0 to 1.

Equation 2.9: 𝐶𝑄 = 𝑒 2 𝑉𝑧𝑧 𝑄/ℎ

Equation 2.10: 𝜂 = (𝑉𝑥𝑥 − 𝑉𝑦𝑦 )/𝑉𝑧𝑧

The relationship between charge distribution symmetry and CQ is most readily illustrated

in Figure 2.1, showing increasing asymmetry resulting in larger CQ values.

22
Figure 2.1. Illustration of 27Al CQ values (in MHz) with varying degrees of symmetry and coordination
numbers. Taken from Kentgens.90

According to perturbation theory and Hamiltonian theory, it is typically the first-

and second-order quadrupolar interactions which are accounted for given by the

Hamiltonians: T̂

(1) 𝑞𝑄𝑒 2
Equation 2.11: Ĥ𝑄 = √6 𝑉 𝑓𝑄
4𝐼(2𝐼−1) 20 20

23
2
(2) 𝑞𝑄𝑒 2 1 2 𝑄 𝑄
Equation 2.12: Ĥ𝑄 = − (4𝐼(2𝐼−1)) 𝜔𝐿 5
𝑋 {(−3√10𝑇̂̂30 + 𝑇̂̂10 (3 − 4𝐼(𝐼 + 1))) 𝑉00 +

(−12√10𝑇̂̂𝑄 𝑄 𝑄
30 − 𝑇̂̂10 (3 − 4𝐼(𝐼 + 1))) 𝑉20 + (−34√10𝑇̂̂30 +

𝑄
3𝑇̂̂10 (3 − 4𝐼(𝐼 + 1)))𝑉40 }

𝑄
where 𝑇̂̂𝑘0 terms are spherical tensors defined by:

𝑄
Equation 2.13: 𝑇̂̂10 = 𝐼𝑍

1
Equation 2.14: 𝑇̂̂𝑄 2
20 = √6 (3𝐼𝑍 − 𝐼(𝐼 + 1))

1
Equation 2.15: 𝑇̂̂𝑄 2
30 = √10 (5𝐼𝑍 − 3𝐼(𝐼 + 1) + 1)𝐼𝑍

Geometrical terms are contained in Vk0:

𝑘
Equation 2.16: 𝑉𝑘0 = ∑𝑛 𝐷𝑛0 (𝛼, 𝛽, 𝛾)𝐴𝑘𝑛

𝑘 (𝛼,
Equation 2.17: 𝐷𝑛0 𝛽, 𝛾) = 𝑒𝑥𝑝(−𝑖𝛼𝑞 ′ ) 𝑒𝑥𝑝(−𝑖𝛾𝑞)𝑑𝑞′𝑞
𝑘
(𝛽)

𝑘
Where 𝐷𝑛0 (𝛼, 𝛽, 𝛾) are Wigner rotational matrices through which the EFG tensor is

transformed from its PAS frame to the laboratory frame, (𝛼, 𝛽, 𝛾) are the Euler

angles associated with such rotation as illustrated in Figure 2.2, and dkq′q (β) is a

reduced Wigner function. Akn terms are given by:

24
1
Equation 2.18: 𝐴00 = − (3 + 2𝑄 )
5

1 1 3
Equation 2.19: 𝐴20 = 14 (2𝑄 − 3), 𝐴2±2 = 7 √2 𝑄

1 3 5 1
Equation 2.20: 𝐴40 = 140 (18 + 2𝑄 ), 𝐴4±2 = 70 √2 𝑄 , 𝐴4±4 = 4 2
√70 𝑄

Figure 2.2. Euler angles defining the direction of B 0 in the principal axis system (PAS) of the EFG
under static conditions.

The first-order quadrupole effect is independent of the Larmor frequency, ωL,

meaning that the central transition (m = +1/2 ↔ -1/2) and symmetric multiple quantum

transitions (i.e. +3/2 ↔ -3/2, +5/2 ↔ -5/2, etc…) are unaffected by the strength of the

magnetic field. The second-order quadrupole effect, however, is inversely proportional to

ωL such that the effects of quadrupolar coupling are decreased with increasing magnetic

field strength. Using the above equations, the transition frequencies from the quadrupolar

coupling associated with first and second-order contributions can be given by:

(1) 3𝑒 2 𝑞𝑄
Equation 2.21: ∆𝜔𝑄 = 4𝐼(2𝐼−1) (2𝑚 + 1)𝑉20

25
2
(2) 𝑒 2 𝑞𝑄 2 2
Equation 2.22: 𝛥𝜔𝑄 = − (4𝐼(2𝐼−1)) 𝜔𝐿 5
∗ {(𝐼(𝐼 + 1) − 9𝑚(𝑚 + 1) − 3)𝑉00 +

(8𝐼(𝐼 + 1) − 36𝑚(𝑚 + 1) − 15)𝑉20 + 3(6𝐼(𝐼 + 1) −

34𝑚(𝑚 + 1) − 13)𝑉40 }

As will discussed in further detail, the quadrupolar nucleus primarily studied in this
93
work is spin I=9/2 Nb. As shown in Figure 2.3, in a B0 magnetic field, there are ten

(2I+1) Zeeman energy levels for the 93Nb nucleus with different precession frequencies at

equilibrium magnetization. Transitions between consecutive energy states are referred to

as single-quantum (1Q) transitions while transitions between non-consecutive states are

referred to as multiple-quantum (MQ) transitions. The transition between Zeeman

eigenstates +1/2 and -1/2 is the central transition (CT) while any other single quantum

9 7 7 5 5
transitions between other eigenstates such as (± 2 ↔ ± 2) , (± 2 ↔ ± 2) , (± 2 ↔

3 3 1
± 2) , 𝑎𝑛𝑑 (± 2 ↔ ± 2) are considered satellite transitions (STs). An example of peaks

due to the CT and ST’s is shown in Figure 2.4a where satellite transitions cover a much

wider frequency range and have substantially lower intensities than the CT. It should be

noted that often if ST peaks are not well resolved they manifest as a broad baseline

overlapping with the CT which can be easily subtracted from the resulting powder pattern.

The powder pattern is the result of lines from different molecular orientations overlapping

over a range of frequencies thus forming a continuous lineshape. As the above equations

indicate, the CT frequency is unaffected by first-order quadrupolar effects. The CT

26
lineshape under static conditions also exhibits easily observable characteristic appearances

depending on the asymmetry parameter as seen in Figure 2.4b. This significantly improves

the ease of extracting the EFG tensor values under certain experimental conditions

(discussed further in subsequent sections) because the first term in the second-order

quadrupolar interactions has no PAS orientation dependence.

Figure 2.3. Energy level diagram91 of spin I=9/2 nucleus and first- and second-order quadrupolar
contribution where ωL is the Larmor frequency.

27
a)

b)

Figure 2.4. a) Simulated powder pattern using only first-order quadrupolar effects displaying all
transitions for an I=9/2 nucleus with CQ = 20 MHz and η = 0 at a magnetic field strength of 9.4 T. b)
Simulated powder patterns using both first- and second-order effects of the central transition for
CQ = 20MHz with η = 0, 0.5, and 1.0.92

Other interactions related to the local bonding electrons result in the

chemical shift anisotropy (CSA), which is dependent on the local symmetry of the metal

28
site.86 The CSA tensor is described by three diagonal elements in its respective PAS, δ11,

δ22, and δ33. Three relevant parameters arise from combinations of the tensor components,

the isotropic chemical shift, δiso, the span (related to the magnitude of chemical shift

anisotropy), Ω, and the skew (related to the symmetry of CSA), κ, defined in Equations

2.23-2.25 using the Herzfeld-Berger Convention.93 As llustrated in Figure 2.5, δiso is

commonly regarded as the center of gravity for the spectrum while Ω describes the

magnitude of the anisotropy. The skew describes the asymmetry of the chemical shift

tensor with values for the skew ranging from -1 to 1.86 When the skew is equal to +1 or -

1 then δ22 must be equal to δ11 or δ33 thus implying the chemical shift tensor exhibits axial

symmetry. This serves as a quick test as to the veracity of the crystal structure being used.

As will be discussed in later chapters, departure of the skew from axial symmetry is

indicative of presence of an orthorhombic space group as opposed to a tetragonal space

group. A set of Euler angles (see Figure 2.1), α, β, and γ are used to define the relative

orientation of one tensor in terms of another.

Equation 2.23: 𝛺 = 𝛿11 − 𝛿33

Equation 2.24: 𝜅 = 3(𝛿𝑖𝑠𝑜 − 𝛿22 )/𝛺

0
Equation 2.25: 𝛿𝑖𝑠 = (𝛿11 + 𝛿22 + 𝛿33 )/3

29
Figure 2.5. Parameters of the Herzfeld-Berger convention as they relate to the NMR spectrum.

2.3 93Nb Quadrupolar Interactions

The NbO6 octahedra are key components of the layered perovskites studied in this
93
work and it has been found that the EFG of the Nb nuclei provide incredibly useful

information regarding the local octahedral structures. 93


Nb has a gyromagnetic ratio of γ

= 6.56 x 107 rad/sT on par with the more commonly studied 13C nucleus, but with 100%

natural abundance making it a desirable nucleus to study. It does however, have a spin of

I = 9/2 and a substantial quadrupole moment Q = -32 x 10-32 m2 resulting in large second-

order quadrupolar contributions to the lineshape of the CT spectrum. As will become

abundantly clear, 93Nb spectra are dominated by second-order quadrupolar coupling which

can be >100 MHz for Nb(V) complexes resulting in signals spanning several hundred kHz.

30
The parameters of the EFG tensor for 93Nb in layered perovskites is dependent on

the octahedral structure with quadrupolar coupling typically found in the wide range of 10-

120 MHz and isotropic chemical shift values from -650 ppm to -1600 ppm, as shown in

Figure 2.6, depending on the coordination of the surrounding oxygen environments.94


93
Traditionally, the analysis of Nb NMR spectra has been limited to highly crytalline

phases, but in recent years, novel techniques have been developed and optimized such that

the highly disordered materials studied in this work, containing multiple overlapping 93Nb

signals, can be successfully characterized.

NbO4

NbO5

NbO6

NbO7

NbO8

500 750 1000 1250 1500 1750 2000


δiso, ppm

Figure 2.6. 93Nb NMR chemical shifts for NbOx polyhedral.94

2.4 Variable Offset Cumulative Spectra (VOCS)

Collecting static 93Nb spectra is done through the use of the Hahn Echo sequence95

rather than a single pulse. This allows for refocusing of magnetization lost in the dead time

and distorted by ringing signals from the probe. Unfortunately, the static 93Nb spectra with

large quadrupolar coupling can span up to several MHz making it impossible to irradiate

31
uniformly at one time using the standard hard pulse. Massiot96-97 developed the VOCS

technique as a method of collecting the full static powder pattern without spectral

distortion. The VOCS method requires the collection of a series of Hahn Echo spectra with

gradually changing offsets such that the final spectrum can be reconstructed by the

summation of the series as illustrated in Figure 2.7.

Figure 2.7. 93Nb VOCS spectra for RbSr2Nb3O10 at 14.1 T with offsets of 62.5 kHz between neighboring
slices.98 The summation of the slices is given with and without baseline correction.

32
The power of the rf pulses used in the Hahn echo sequence allows for the

calculation of the irradiation profiles for each echo experiment. From this, the frequency

offset needed can be calculated such that the total irradiation profile is even across the

whole spectrum. The major drawback of the VOCS method is the long experiment time

due to the collection of many spectra (sometimes upwards of 20 individual spectra) making

it an impractical method for the analysis of a large series of samples.

2.5 Magic Angle Spinning

When quadrupolar nuclei are put into an applied magnetic field Bo, they experience

chemical shift anisotropy (CSA), dipolar coupling and quadrupolar coupling.99 These three

main interactions give rise to static spectra with incredibly wide powder patterns spanning

several hundreds of kilohertz (kHz).96, 100-101


Even in ideal samples with only one

quadrupolar environment present, this breadth produces difficulty regarding extracting the

CSA and EFG parameters because the static signal is affected by all three interactions.

Multiple quadrupolar environments, a characteristic of the compounds studied in this work,

further complicates matters making it nearly impossible to extract the physically correct

CSA and EFG parameters rather than simply a mathematically correct fit. A way to

compensate for these complex interactions is to use magic angle spinning (MAS)

techniques.102-105

In an MAS experiment, the sample is spun about an axis inclined at an angle θR to

the applied field B0. The CSA contribution to spectral frequency, for axially symmetric

systems, is given by:

33
𝑎𝑛𝑖𝑠𝑜 (𝜃) 𝑃𝐴𝐹 1
Equation 2.26: 𝜔𝐶𝑆 = −𝜔0 (𝜎𝑍𝑍 − 𝜎𝑖𝑠𝑜 ) 2 (3𝑐𝑜𝑠 2 𝜃 − 1)

while the heteronuclear dipolar coupling Hamiltonian is given by:

𝜇 𝛾𝐼 𝛾𝑆 ħ
̂𝑑𝑑
Equation 2.27: 𝐻 ℎ𝑒𝑡𝑒𝑟𝑜
= − (4𝜋0 ) 𝑟3 𝑍 𝑍
𝐼̂ 𝑆̂ (3𝑐𝑜𝑠 2 𝜃 − 1)

When considering the relationship between θ and θR through the expression:

1
Equation 2.28: 〈3𝑐𝑜𝑠 2 𝜃 − 1〉 = 2 (3𝑐𝑜𝑠 2 𝜃𝑅 − 1)(3𝑐𝑜𝑠 2 𝛽 − 1)

where β is the angle between the principle z-axis of the chemical shielding tensor and the

spinning axis, setting θR to 54.74o and spinning at a fast-enough rate brings the (3𝑐𝑜𝑠 2 𝜃𝑅 −

1) term to zero, thus eliminating the effects of CSA and dipolar coupling in the MAS

spectrum.

When considering the quadrupolar contributions to the spectral signal, the V20

terms in the first- and second-order contributions will be averaged to zero. However, a

large enough quadrupolar interaction cannot be completely averaged out under fast MAS

(typically 20-30 kHz in the work performed here) thus resulting in the presence of spinning

sidebands. While the CT will not be effected by MAS to the first order, STs will be broken

into low intensity spinning sidebands. Second-order broadening effects can only be

partially reduced under MAS since the V40 term does not average to a nonzero value.

Figure 2.8 illustrates both the benefits and limitations of MAS alone. While a

spectrum for a 93Nb nucleus with CQ = 20 MHz (on the lower end of typical couplings seen

in this work) does exhibit substantial reduction in CSA contributions at the maximum

34
spinning speeds attainable with our instrumentation, they are not completely removed and

thus the spectrum is not highly resolved. Spinning at much higher speeds approaching 100

kHz does produce high resolution spectra, but unfortunately these spinning speeds are by

no means routine and in fact only a handful of solid-state NMR laboratories can perform

such experiments.106-109 Therefore, more specialized MAS based experiments, to be

discussed further, are required to accurately isolate and extract the different EFG

parameters without producing a mathematical fit that is not necessarily physically correct.

Figure 2.8. 93Nb static and MAS spectra at 9.4 T with varying spin rates (C Q = 20 MHz, η = 0.2, δiso =
0ppm, Ω = 500 ppm, κ = 0.5, α = β = γ = 0). Spectra were simulated using Simpson software.110

35
2.6 Wideline Uniform Rate Smooth Truncations (WURST) Pulses

The use of the VOCS method for obtaining wideline NMR spectra, while effective,

is substantially inefficient considering the time needed to collect a full spectrum. The

VOCS method can obtain slices of the full spectrum in increments of only 72.5 kHz

meaning a signal spanning 700 kHz (a range very common in the compounds studied in

this work) would require ten spectra be obtained. Further complicating matters, if the range

of the signal coverage is unknown (again, a very common occurrence in our compounds)

then slices extending substantially outside of the signal range are required to determine if

the full spectrum has been obtained. WURST broadband excitation pulses111-116 overcome

this limitation by drastically increasing the excitation bandwidth to the point that a 700 kHz

spectrum can be obtained in a single experiment. It should be noted that uniform excitation

is only when the excitation bandwidth is limited to 300 kHz using our instrumentation, but

the use of broader excitation profiles is still useful in determining the range of signal

present.

The time-dependent amplitude profile for WURST pulses is given by the equation:

𝜋𝑡
Equation 2.29: 𝜔1 (𝑡) = 𝜔𝑚𝑎𝑥 (1−∣ 𝑐𝑜𝑠 (𝜏 ) ∣𝑁 )
𝑤

Where ωmax is the maximum RF amplitude, τw is the pulse duration, and N determines the

extent to which the edges of the pulse are rounded off. As can be seen in Figure 2.9, higher

N values give more rectangular amplitude profiles.

36
Amplitude WURST-N Amplitude Profiles

Time

WURST-2 WURST-4 WURST-40 WURST-80

Figure 2.9. WURST-N Amplitude profiles for WURST-2, WURST-4, WURST-40, and WURST-80
pulses.

Kupče and Freeman originally designed WURST pulses for adiabatic inversion of

magnetization.117 Adiabatic NMR experiments require pulses satisfying the adiabatic

condition (|ωeff(t)| >> |dα/dt|, where dα/dt is the instantaneous angular velocity) such that

magnetization will rotate at a constant flip angle, gradually following Beff during the

frequency sweep.118 It has been found that under nonadiabatic conditions comparable to

those employed in this work, WURST pulses can still be utilized in obtaining wideline

NMR spectra of quadrupolar nuclei.119 In the non-adiabatic regime, the WURST pulse can

be treated as a train of π/2 pulses whose individual frequencies vary linearly with time and

thus excite the signal with matching frequency, but having no effect on others. Signal

excitation then becomes time-dependent and the total spectrum will develop a frequency-

dependent phase difference, Φ(ν), given by the equation:119

3𝜏𝑒𝑥𝑐 𝜏
Equation 2.30: 𝛷(𝜈) = ( 2
)𝜈 𝑒𝑥𝑐
− ( 2𝛥 ) 𝜈2

37
Furthermore, since acquisition does not begin until the end of the pulse, signal excited at

the beginning of the pulse will be more attenuated due to dephasing and T2 relaxation than

signal excited at the end of the pulse as shown in Figure 2.10a. To overcome this problem

of phasing, application of a second WURST pulse to refocus magnetization can be used.

Assuming the refocusing pulse has an identical sweep range, Δ, the time-dependent phase

distribution then becomes:

𝜏𝑒𝑥𝑐 𝜏𝑒𝑥𝑐 𝜏𝑟𝑒𝑓


Equation 2.31: 𝛷(𝜈, 𝑡) = (𝑡 − − 𝜏𝑑 ) 𝜈 + ( − ) 𝜈2
2 2𝛥 𝛥

where τd is the delay between pulses. Therefore, if τref is set such that 2τref = τexc then all

signals will refocus at the same point in time thus negating the second order phase

distortions (Figure 2.10b,c).

Figure 2.10. Schematic illustrations of WURST excitation and refocusing pulses reproduced from
Schurko et al.115 A - Single excitation pulse showing signal attenuation during the excitation pulse. B –
A refocusing pulse with 2τref = τexc leads to removal of the second-order phase distortion. C – A
refocusing pulse with τref = τexc does not eliminate second-order phase distortions as illustrated by the
displaced echoes shown in blue and red.

38
To further improve the efficiency, the WURST echo scheme was developed into

the WURST-QCPMG pulse sequence. As depicted in Figure 2.11, in the quadrupolar

Carr-Purcell-Meiboom-Gill (QCPMG120) pulse sequence, an off shoot of the CPMG

sequence,121 magnetization is refocused and detected N times in the cycle. Despite signal

attenuation due to relaxation as the train of echoes progresses, a substantial increase in

signal to noise is obtained. The use of WURST pulses in the CPMG cycle follows this

trend. The significant S/N improvements and greatly increased excitation bandwidth has

made WURST-QCPMG a desirable pulse sequence for use with quadrupolar nuclei.112-113

Figure 2.11. QCPMG Pulse Sequence.120

Traditionally, the processing of CPMG sequences produces a spikelet pattern after

Fourier transformation (as shown in Figure 2.12b) with equally spaced peaks whose

intensities resemble the conventional powder pattern. To assist with spectrum fitting and

analysis, co-addition of each echo in the CPMG train before Fourier transformation

produces a conventional powder pattern spectrum.

39
Figure 2.12. 91Zr NMR powder patterns obtained from a ZrO2 MAS rotor at 9.4 T using (a) WURST
echo and (b) WCPMG pulse sequences. Values shown on the right are the relative signal intensities,
normalized to account for the differing number of scans. The simulation in (c) includes central
transition lineshapes for the tetragonal (dotted) and orthorhombic (dashed) phases. Illustration was
obtained from O’Dell.116

Recently, the suite of WURST related pulse sequences has grown to include the

Broadband Adiabatic Inversion Cross-Polarization122 (BRAIN-CP) sequence shown in

Figure 2.13. Cross Polarization (CP) is used extensively in solid-state NMR as a method

40
of signal boosting where polarization is transferred from a higher frequency nucleus

(commonly 1H) to a lower frequency nucleus (such as 13


C).123-125 This transfer can be

achieved by spin locking the transverse 1H magnetization and irradiating the low frequency

nucleus during the contact period such that the Hartman-Hahn match,123 ν1(1H) = ν1(X), is

satisfied. Unfortunately, when dealing with large anisotropic interactions, CP using

traditional spin-locking fields is limited due to the bandwidth being determined by the

strength of the spin-locking fields and thus only a small fraction of the total powder pattern

is excited.

BRAIN-CP overcomes this excitation limitation using a standard pulse to maintain

the 1H spin lock, but with a WURST pulse on the X channel sweeping magnetization

adiabatically from orientation along the +Z axis to along the –Z axis over a wide

bandwidth. Under these sweeping conditions, each isochromat meets the Hartmann-Hahn

matching condition at different points in time during the sweep. CP can then be uniformly

performed over a broad range of frequencies with hardware considerations being the

limiting factor. The time interval during which CP occurs when ω1,I = ω1,S for each

isochromat can be estimated using the equation:

3 24𝜋𝜔1,𝑆
Equation 2.32: 𝜏𝐶𝑃 = √
𝑅2

where R is the sweep rate. Under a typical experiment with a 10 ms, 20 kHz WURST

pulse sweeping over 150 kHz, the approximate cross polarization time for each isochromat

comes to be 1.0 ms. A final π/2 pulse orients the magnetization into the x-y plane (a

necessary condition for signal observation).

41
Figure 2.13. BRAIN-CP pulse sequence.

2.7 Multiple-Quantum Magic Angle Spinning (MQMAS)

As discussed above, it is possible in principle to remove anisotropic broadening

under sufficiently fast MAS conditions, but in practice it is often impossible to reach speeds

sufficiently fast to completely remove such interactions. Take for instance a case where

the quadrupolar interaction is very strong (a recurring motif for samples studied in this

work). While MAS helps to narrow the lineshapes, the second order broadening is too

strong to be completely averaged to zero by MAS alone. Fortunately, this is by no means

an insurmountable obstacle. Various high-resolution MAS based NMR techniques have

been created to further reduce or separate anisotropic interactions. Dynamic Angle

Spinning (DAS126) and Double Rotation (DOR127) are two early developments, but both

are encumbered by mechanical complexities (see Figure 2.14). Both methods operate

through spinning about two different axes with DAS requiring switching of angles and

DOR requiring simultaneous spinning of two rotors positioned inside one another with

each spinning about a separate axis. The technique which has really taken off and become

42
a mainstay of solid-state NMR is MQMAS, first created in 1995 by Frydman and

Harwood.128

Figure 2.14. Schematic representation of the DOR sample setup requiring the spinning about two axes
simultaneously.

Under MAS conditions, the second-order quadrupolar frequency term is given by:

(2)
Equation 2.33: 𝜈𝑄 (𝑚, 𝛽) = 𝜈0𝑄 𝐶0𝐼 (𝑚) + 𝜈2𝑄 (𝜃, 𝜑)𝐶2𝐼 (𝑚)𝑃2 (𝑐𝑜𝑠𝛽) +

𝜈4𝑄 (𝜃, 𝜑)𝐶4𝐼 (𝑚)𝑃4 (𝑐𝑜𝑠𝛽)

where {νQl}l=0,2,4 corresponds to the isotropic, second, and fourth-rank quadrupolar

contributions, respectively. {CIl(m)}l=0,2,4 are polynomials which depend on spin I and

coherence order m.

Equation 2.34: 𝐶0𝐼 (𝑚) = 2𝑚[𝐼(𝐼 + 1) − 3𝑚2 ]

43
Equation 2.35: 𝐶2𝐼 (𝑚) = 2𝑚[8𝐼(𝐼 + 1) − 12𝑚2 − 3]

Equation 2.36: 𝐶4𝐼 (𝑚) = 2𝑚[18𝐼(𝐼 + 1) − 34𝑚2 − 5]

P2(cosβ) and P4(cosβ) terms are the second and fourth-rank Legèndre polynomials given

by

3𝑐𝑜𝑠2 𝛽−1
Equation 2.37: 𝑃2 (𝑐𝑜𝑠𝛽) =
2

35𝑐𝑜𝑠4 −30𝑐𝑜𝑠2 𝛽+3


Equation 2.38: 𝑃4 (𝑐𝑜𝑠𝛽) = 8

At the magic angle, 3cos2β-1 equals zero, and thus the second-rank term disappears while

the fourth-rank term remains. The second-order quadrupolar shift for the central line in the

fast rotation regime is now given by:

(2)𝑀𝐴𝑆 1 3𝜒 2 3
Equation 2.39: 𝜔 11 = − [ ] [𝐼(𝐼 + 1) − ] 𝑥 [𝐷(𝛼, 𝜂)𝑐𝑜𝑠 4 𝛽 +
− , 6𝜔0 2𝐼(2𝐼−1) 4
22

𝐸(𝛼, 𝜂)𝑐𝑜𝑠 2 𝛽 + 𝐹(𝛼, 𝜂)]

with

21 7 7
𝐷(𝛼, 𝜂) = − 𝜂𝑐𝑜𝑠2𝛼 + (𝜂𝑐𝑜𝑠2𝛼)2
16 8 48

9 1 7
Equation 2.40: 𝐸(𝛼, 𝜂) = − 8 − 12 𝜂 2 + 𝜂 𝑐𝑜𝑠 2𝛼 − 24 (𝜂𝑐𝑜𝑠2𝛼)2

5 1 7
𝐹(𝛼, 𝜂) = − 𝜂𝑐𝑜𝑠2𝛼 + (𝜂𝑐𝑜𝑠2𝛼)2
16 8 48

44
Hence, MAS conditions alone only partially reduce the second-order quadrupolar

interaction. Therefore, to accurately study nuclei containing large quadrupolar

interactions, other experiments such as MQMAS are employed.

The basic structure of the MQMAS experiment involves excitation of unobservable

multiple-quantum coherences (3Q, 5Q, 7Q, or 9Q for I = 9/2 spins) followed by conversion

to the observable single-quantum coherence which allows for processing of the resulting

NMR signals such that interactions can be separated. While a variety of MQMAS pulse

sequences such as two pulse,128 z-filter,129 split-t1,130 and soft pulse added mixing131

(SPAM) have been developed, the research discussed in this work utilizes the shifted-

echo132-134 version of the 3QMAS pulse sequence. The 3QMAS pulse sequence is shown

in Figure 2.15, with an initial pulse generating the +3 coherence, followed by a FAM-II135-
137
train of conversion pulses generating the +1 coherence, and finally followed by a π pulse

to select the central transition thus generating a shifted echo such that the whole echo can

be collected.

45
Figure 2.15. Shifted-echo 3QMAS sequence with a FAM-II train of conversion pulses and the
corresponding coherence transfer pathway.

Processing of the 3QMAS data was performed using a double shearing method32

that successfully separates interactions such that only the quadrupolar chemical shift

(containing information about the EFG tensor) remains in the F1 dimension while the F2

dimension consists of both isotropic chemical shift and quadrupolar shift interactions. The

frequencies in each dimension following the double shearing can be described as:

Equation 2.41: 𝜈1 = 𝜈0𝑄 𝑘𝑄1 + 𝜈4𝑄 (𝜃, 𝜑)𝑃4 (𝑐𝑜𝑠𝛽)𝐶41 (−1)

10
Equation 2.42: 𝜈2 = 𝜈0𝐶𝑆 − 27 〈𝜈1 〉 + 𝑛𝑣𝑅

where 𝜈0𝐶𝑆 is the isotropic chemical shielding, n is an integer, vR is the sample spinning

frequency, and

46
2
𝐶𝑄 (3+𝜂2 )
Equation 2.43: 𝜈0𝑄 = − 2
10𝜈𝐿 (2𝐼(2𝐼−1))

648
Equation 2.44: 𝑘𝑄1 = 𝑓𝑜𝑟 𝐼 = 9/2
17

and since 𝜈4𝑄 (𝜃, 𝜑) ≈ 0 the first moment of the powder pattern simplifies to:

Equation 2.45: 〈𝜈1 〉 = 𝜈0𝑄 𝑘𝑄1

The quadrupolar coupling (CQ) and asymmetry (η) parameters can then be readily extracted

from the resulting sheared spectrum. Due to this shearing, the dwell time in the indirect
1
dimension must be scaled by − 𝜆(𝐼,𝑝) where 𝜆(𝐼, 𝑝) = -72/34 for I=9/2 and the coherence

order p = +3

Fernandez et al.138 quite cleverly combined the MQMAS experiment with S-

observe, I-dephase Rotational Echo Double Resonance (REDOR), herein referred to as

MQMAS-t2-REDOR (Figure 2.16), to create a new tool for the study of internuclear

distances in complex systems. The REDOR sequence81-82 utilizes rotor-synchronized rf

pulses to reintroduce heteronuclear dipolar coupling between I spins and the observed S

spins resulting in signal attenuation for the observed S spins. As will be discussed further

in Chapter 5, this allows for easy assignment of 93Nb signals to either interior or exterior
93
NbO6 octahedra in the layered structure since the Nb-1H internuclear distances for the

interior and exterior NbO6 octahedra are such that only signal due to exterior NbO6

octahedra will undergo observable attenuation.

47
Figure 2.16. MQMAS-t2-REDOR pulse sequence. The I and S channels refer to the 1H and 93
Nb
channels, respectively. An R3 recoupling sequence was used.139-140

As the following chapters will show, these SSNMR techniques have been

invaluable in the characterization of the layered materials studied in this work. Without

SSNMR, a substantial portion of the conclusions drawn from this work would lack the

substantial experimental evidence needed to be accepted as accurate and should serve as a

great example of the potential for using SSNMR in the study of layered perovskites.

48
Chapter 3: Experimental Overview

3.1 Materials

Raw materials, RbNO3, K2CO3, CaCO3, SrCO3, and Ba(NO3)2 salts were purchased

from Sigma-Aldrich and used without further purification. RbCl and Nb2O5 were

purchased from Alfa Aesar and used without further purification. Alcohols were purchased

from Alfa Aesar and diluted with distilled water. 6M HNO3(aq) was diluted from

concentrated HNO3 using distilled water. Grafting reactions were performed using a Parr

23 mL microwave acid digestion bomb and a 1200 W domestic microwave oven.

3.2 Synthetic Procedures*

3.2.1 Three-layer Niobates Using MSS

RbSrxCa2-xNb3O10 and RbSrxBa2-xNb3O10 samples were synthesized using molten

salt heating31, 141-142 of the carbonate and oxide precursors in excess RbCl (the salt in molten

salt synthesis). K2CO3, SrCO3/CaCO3/Ba(NO3)2, and Nb2O5 were added in a 1.25:2:3 ratio

of K : Sr/Ca/Ba : Nb and ground together for five minutes before light mixing with 20 fold

molar excess RbCl. K2CO3 was added in a slight excess to account for its volatile nature

(or rather the volatile nature of K2O produced during the synthesis143). The samples were

then heated at 800-900 oC for 30-45 minutes followed by washing with warm distilled

water to remove the excess RbCl. The acid exchanged samples were obtained by mixing

the parent Rb forms with 6 M HNO3 for 3 days at 50 oC. As a rule of thumb, 1g of sample

is treated with ~75 mL of 6 M HNO3.

*The procedures presented in this chapter are generalized. Details for individual samples are presented in
subsequent chapters.

49
3.2.2 Three-layer Niobates Using Microwave Heating

Some RbSrxCa2-xNb3O10 and RbSrxBa2-xNb3O10 samples were prepared using

microwave synthesis.144 Stoichiometric amounts of Rb2CO3, CaCO3/SrCO3/Ba(NO3)2,

and Nb2O5 were ground together and placed in a thermally insulated box containing silicon

carbide microwave susceptors. The box was then heated using an 1100 W domestic

microwave oven for 5-10 min on high power followed by 10-15 minutes at 50% power to

reach an internal temperature of 1000-1150 oC. The samples were then washed with

distilled water to remove excess Rb2CO3 and dried at 110 oC. The usual acid exchange

procedure was employed to obtain the protonated forms.

3.2.3 Alkyl Grafting with Conventional Heating

C1/Ca2Nb3O10 was synthesized by mixing ~0.2 g HCa2Nb3O10 with 12.4 mL

methanol and 1.1 mL distilled water for a 90% w/w methanol solution. The sample was

then sealed in a general-purpose acid digestion bomb and heated at 100 oC for 3 days

followed by filtration and washing with acetone. C3/Ca2Nb3O10 was synthesized by

mixing the methoxylated sample with 9.1 mL n-propanol and 0.4 mL distilled water for a

95% w/w n-propanol solution followed by heating in an acid digestion bomb at 150 oC for

7 days followed by filtration and washing with acetone. C6/Ca2Nb3O10 was synthesized

by mixing the C3 sample with 10.0 mL n-hexanol followed by heating in an acid digestion

bomb at 150 oC for 7 days followed by filtration and washing with acetone.

50
3.2.4 Alkyl Grafting with Microwave Heating

C1/Sr2Nb3O10 was synthesized by mixing ~0.2 g HSr2Nb3O10 with 12.4 mL

methanol and 1.1 mL distilled water for a 90% w/w methanol solution. The sample was

then sealed in a microwave acid digestion bomb and heated in a domestic 1200W

microwave oven using a cycled heating scheme to prevent overheating of the acid digestion

bomb. The sample was heated at 100% power for 20s followed by a 40s cooling period

repeated five times in succession. The sample was left to cool for 15-20 min, then the

process was repeated a total of ten times. The sample was then filtered and washed with

acetone. C3/Sr2Nb3O10 was synthesized by mixing the methoxylated sample with 9.1 mL

n-propanol and 0.4 mL distilled water for a 95% w/w n-propanol solution followed by

heating at 100% power for 40s followed by a 20s cooling period repeated five times in

succession. The sample was left to cool for 15-20 min, then the process was repeated a

total of ten times. The sample was then filtered and washed with acetone. C6/Sr2Nb3O10

was synthesized by mixing the C3 sample with 10.0 mL n-hexanol followed by heating at

100% power for 60s followed by a 15s cooling period repeated four times in succession.

The sample was left to cool for 20-25 min, then the process was repeated a total of six

times. The sample was then filtered and washed with acetone. The microwave heating

reaction cycles are shown in Table 3.1.

51
Table 3.1. General grafting procedure for Ca xSr2-xNb3O10 compounds.
Alcohol Heating Cycle (100% # of Cycles Cooling Period Total Reaction
power) Between Cycles Time
Methanol 5x (20s on 40s off) 10 15-20 min ~4 hours
(microwave)
Methanol 100o C 3 days --- --- 3 days
(conventional)
Propanol 5x (40s on 20s off) 10 15-20 min ~4 hours
(microwave)
Propanol 150o C 7 days --- --- 7 days
(conventional)
Hexanol (microwave) 4x (60s on 15s off) 76 20-25 min ~3.5 hours
Hexanol 150o C 7 days --- --- 7 days
(conventional)

3.2.5 RbCa2NaNb4-xTaxO13 and HCa2NaNb4-xTaxO13

RbCa2Nb3O10 was synthesized using the molten salt technique described

previously. NaNbO3 and NaTaO3 were synthesized using the molten salt heating method

as well. Na2CO3 and Nb2O5 (for NaNbO3) or Ta2O5 (for NaTaO3) were ground together in

a 1:1 ratio for Na:Nb/Ta. The sample was then mixed with 10 molar excess of a

K2SO4/Na2SO4 1:1 mixture and heated at 825 oC for 15 minutes. It was then washed with

warm distilled water to remove excess K2SO4/Na2SO4. It should be noted that the 15-

minute reaction time was used to ensure small particle formation, which is essential to the

four-layer synthesis

RbCa2Nb3O10 and NaNbO3 or NaTaO3 were ground together for 5 minutes in a 1:3

molar ratio followed by rapidly heating in a thermally insulated box containing silicon

carbide microwave susceptors at 1200 oC for 20 minutes in a ThermWave microwave oven

with a platinum thermocouple from Research Microwave Systems, LLC. To obtain the

protonated form of the compound, the Rb parent compound was mixed with 6M HNO3 and

heated at 80 oC for 3 days with acid replacement every day.

52
3.2.6 Nanosheet synthesis using TBAOH

HCa2Nb3O10 and HSr2Nb3O10 exfoliation into nanosheets was performed using the

traditional method of intercalation of TBA+ using a TBAOH aqueous solution. Typically,

a 10:1 molar ratio of TBA+:Nb was employed using a 0.5 wt % TBAOH solution. The

mixture was then gently stirred at room temperature for 7 days. The mixture was then

centrifuged at 6000 rpm for 45 minutes to separate the unexfoliated sediment and the

colloidal suspension consisting of nanosheets. Adding a 0.1 M HNO3 solution dropwise

over the course of several hours until the pH of the solution reaches ~6-7 results in the

flocculation of nanosheets. Freezing of the flocculent followed by exposure to vacuum (~

0.1 Torr) generates solid high surface area nanosheets (though it should be noted the

presence of stacking reflections in the XRD powder patterns indicates some restacking has

occurred).

3.2.7 Nanosheet Synthesis Using Microwave Assisted Grafting

Approximately 0.1g of the alcohol grafted samples was mixed with either 2-

pentanone (2P) or 5-methyl-2-hexanone (5M2H in a glass vial and sonicated in a VWR

Symphony Ultrasonic Cleaner with an operating frequency of 35 kHz for four hours with

the exchange of the water in the sonication bath every 30 minutes to prevent overheating.

The colloidal suspension generated was then sealed in a microwave acid digestion bomb

and heated over a period of approximately four hours using cycled heating similar to that

employed in the grafting procedure with heating cycles of 10x[5x(45s on 15s off)] with 15-

53
20 minute cooling intervals between 5-minute heating cycles. Following the microwave

heating the sample was again placed in a glass vial and sonicated for four hours. The

solutions were then centrifuged at 6000 rpm for 45 minutes and the supernatants containing

the nanosheets were collected. Methods of collecting nanosheets out of solution will be

discussed in detail in Chapter 7.

3.3 Characterization Methods

3.3.1 Powder XRD

All XRD powder patterns were collected using a Bruker AXS D8 Focus

diffractometer, using a graphite monochromator and Cu Kα irradiation. Data were

collected in a θ/2θ mode with a scanning 2θ range of 4-60 degrees or 2.5-14 degrees for

short range grafted samples where the first and second stacking reflections of the perovskite

were of greatest importance. Le Bail145 fits were performed on the XRD patterns using the

TOPAS146 software package from Bruker, and using a Chebychev polynomial background,

lattice parameters, particle size, zero corrections, χ2, and the background subtracted R-

factor were calculated.

3.3.2 ATR-IR

IR spectra were collected using a Perkin Elmer Infrared Spectrophotometer

(Spectrum 100) with an attached Pike Technologies GladiATR accessory featuring a

monolithic diamond crystal. Scans were performed from 450-4000 cm-1.

54
3.3.3 SEM and EDX

A table-top scanning electron microscope (SEM) TM-3000 coupled with an

electron dispersive spectrometer (EDX) was used for imaging and elemental analysis. 15

kV was used for all EDX measurement with 300 seconds used for signal averaging. Higher

resolution SEM images were obtained using a Tescan Vega-3 SEM instrument

3.3.4 UV-visible Diffuse Reflectance

UV-visible spectra were collected on a Perkin Elmer Lambda 35 UV/Vis

spectrophotometer with an attached LabSphere RSA-PE-20 diffuse reflectance and

transmittance accessory using a 160 nm/min scan rate from 200 – 1000 nm. Barium sulfate

was used for background collection.

3.3.5 Solid-State NMR Experiments

All 93Nb SSNR data were collected at magnetic field strengths of 9.4 and 14.1 T on

Varian INOVA spectrometers with resonance frequencies of 97.85 and 146.61 MHz,

respectively. All 1H, 23


Na, and 87
Rb NMR experiments were performed at 9.4 T with

resonance frequencies of 399.76, 104.61, and 67.96 MHz, respectively. 2.5 mm, 5.0mm,

and 7.0mm Chemagnetics double resonance MAS probes were used for experiments at 9.4

T. A 5 mm Varian triple resonance liquid state probe was used for all experiments on the

14.1 T instrument.

55
VOCS

VOCS experiments were performed using 15-21 Hahn Echo spectra with offset step

sizes of 78.125 and 62.5 kHz for the 9.4 and 14.1 T instruments, respectively. The π/2

pulse lengths were 2.4 and 4.0 μs at 9.4 and 14.1 T, respectively. The rf field strengths

were 20.8 and 12.5 kHz, respectively.

WURST

Experiments using WURST pulses were carried out at both 9.4 and 14.1 T fields.

Pulse parameters were optimized using a powder sample of RbSr2Nb3O10. The static 93Nb

NMR spectra for RbSr2Nb3O10 were found to have well defined features making it an ideal

test compound for the WURST pulse sequences through the comparison with VOCS

spectra. Using the intensity of one singularity in the RbSr2Nb3O10 spectrum and

incrementally changing the offset of the WURST pulse, the shape vs. offset profile was

optimized and it was found that an excitation bandwidth of 300 kHz gave an optimal

excitation profile. 50 μs pulses were used with 4 kHz and 8 kHz rf field strengths on the

initial excitation pulse and subsequent refocusing pulses, respectively.

Triple-Quantum Magic Angle Spinning (3QMAS)

3QMAS experiments used FAM-II conversion pulses under spinning rates of 25

kHz for 93Nb experiments, 20 kHz for 23Na experiments, and 10 kHz for 1H experiments.

A 1.7 μs π/2 initial excitation pulse was followed by the four-pulse FAM-II sequence with

pulse lengths of 0.6 μs at rf field strengths of 85 kHz. The final π pulse used to convert to

56
the -1 coherence order was 2.0 μs with an rf field strength of 50 kHz. Rotor synchronization

was applied on the anisotropic dimension with 8 to 16 data points depending on the

relaxation of each sample.

Multi-Nuclear Experiments
93
BRAIN-CP experiments were collected at 9.4 T with Nb WURST pulses set to

an rf power of 3 kHz and a sweep width of 300 kHz and 1H pulses set to an rf power of

62.5 kHz. The contact time used in the CP experiments was 10.0 ms with a recycle delay

of 3.0 s. The MQMAS-t2-REDOR sequence utilized 93Nb rf field strengths of 85 kHz with
1
H RF power set to 80 kHz and an R3 recoupling sequence139-140 on the 1H channel.

NMR Data Processing and Spectral Fitting

NMR data were processed using RMN147 and fit using WSOLIDS,148 DMFIT,149

and QUADFIT.150 WURST-QCPMG spectra were obtained through the co-addition of all

individual echoes within the CPMG echo train to produce a single FID signal for

subsequent processing. Whole-echo collection combined with magnitude spectral

transformation overcame the problem of second-order phase distortions. The SIMPSON

NMR110 simulation program was used to optimize several pulse sequences. The alternative

double shearing scheme described in Chapter 2 was used to process all MQMAS spectra.

57
Chapter 4. Exploration of Local Structural Changes in Layered
Niobates due to Compositional Changes at the A-Site.
4.1 Introduction

As the previous chapters have discussed, the composition of A’[An-1BnO3n+1]

layered perovskites leads to a myriad of different structures and properties. The work in

this chapter is focused on the effect of modifying the A site in RbA2Nb3O10 when A is a

mixture of alkaline earth metal pairs: Ca and Sr, or Ba and Sr. The structure and properties

of the end members RbCa2Nb3O1024, 29, 151-154


and RbSr2Nb3O1030, 151, 154-155
have been

extensively studied while the compound RbBa2Nb3O10 cannot be synthesized using readily

available synthetic routes as will be discussed later. As discussed in Chapter 3 and shown

by previous work within our research group,32, 98 NMR is an invaluable tool when studying

the structure of layered niobates as it provides information regarding local structure not

available with the conventional XRD structural characterization. In particular, the study

of the 93Nb nuclei allows for characterization of both exterior and interior NbO6 octahedra.

A thorough search of the literature has found only one study on the structure of mixed

cation RbCaxSr2-xNb3O10 compounds from Geselbracht et al.31 In that work, the presence

of a single phase in XRD patterns was taken as an indication of homogeneous distribution

of Sr and Ca within the perovskite structure. While the work presented here also indicated

long-range homogeneity based on XRD patterns, NMR spectra clearly show a

heterogeneity regarding the distribution of Ca and Sr within local structural ranges.

58
4.2 Experimental

RbSrxCa2-xNb3O10 and RbSrxBa2-xNb3O10 samples were synthesized using the

molten salt or microwave heating methods described in Chapter 3. Specific heating

temperatures and times for each compound are shown in Table 4.1 and 4.2. A comparison

of samples with identical composition, but with different synthetic techniques, showed no

significant changes in structure based on XRD patterns and NMR spectra. The only

noticeable difference between samples was particle size as observed in SEM images.

Samples synthesized with the molten salt method exhibited noticeably smaller particle

dimensions. The molten salt acts as a solvent-like medium during the reaction, thus

allowing for product formation under milder conditions due to the solubility of oxide

particles in the salt. Decreased particle size compared to dry synthesis may be the result

of more numerous points of nucleation growing simultaneously.

59
Table 4.1. Synthetic conditions for RbCaxSr2-xNb3O10 compounds using dry microwave heating or MSS.
Theor. Ca Exper. Ca Synthesis Mass Mass Mass Mass Mass Mass Rb:K:Ca:Sr:Nb Heating Time Temp
Content Content Type RbCl Rb2CO3 K2CO3 CaCO3 SrCO3 Nb2O5 Molar Ratio (OC)
0 0 Microwave 0 0.1913 0 0 0.3727 0.491 1.25:0:0:2.03:2.97 100% 8 min, 995*
Dry 50% 12 min
0.25 0.4 Microwave 0 0.1818 0 0.0322 0.3261 0.4994 1.25:0:0.26:1.75:2. 100% 7 min, 1049*
Dry 98 50% 12 min
0.5 0.9 Microwave 0 0.185 0 0.0616 0.284 0.4991 1.25:0:0.48:1.50:2. 100% 6.5 min, 972*
Dry 93 50% 12 min
0.75 0.7 Microwave 0 0.1841 0 0.0949 0.2319 0.5038 1.25:0:0.74:1.23:2. 100% 6.5 min, 1031*
Dry 97 50% 12 min
1 1.1 Microwave 0 0.1821 0 0.1248 0.1859 0.5058 1.25:0:0.99:1.01:3. 100% 6.5 min, 967*
Dry 02 50% 12 min
1.25 1.3 Microwave 0 0.1866 0 0.1619 0.1453 0.5158 1.25:0:1.25:0.76:3. 100% 6.5 min, 1039*
Dry 00 50% 12 min
1.5 1.5 Microwave 0 0.1833 0 0.1893 0.0916 0.4982 1.25:0:1.49:0.50:2. 100% 5 min, 952*
Dry 95 50% 11 min
1.75 1.9 Microwave 0 0.185 0 0.2219 0.0461 0.5036 1.25:0:1.73:0.22:2. 100% 6 min, 999*
60

Dry 96 50% 11 min


2 2 Microwave 0 0.189 0 0.2496 0 0.491 1.25:0:1.99:0:2.95 100% 7 min, 918*
Dry 50% 5 min
0 0 Molten Salt 3.021 0 0.0851 0 0.3683 0.4978 20.6:1:0:1.96:2.95 45 min 900
3
0.5 0.5 Molten Salt 3.041 0 0.0887 0.0628 0.285 0.5025 19.6:1:0.49:1.50:2. 30 min 900
95
1 1.1 Molten Salt 3.013 0 0.0852 0.1227 0.1843 0.5 19.6:1:1.01:0.98:2. 30 min 900
7 99
1.5 1.7 Molten Salt 3.018 0 0.0847 0.19 0.0906 0.5027 20.4:1:1.55:0.50:3. 30 min 825
2 09
2 2 Molten Salt 3.013 0 0.0865 0.2536 0 0.5008 19.9:1:2.02:0:3.01 25 min 800
7
* Temperature was recorded shortly after removal from the microwave and thus reflects an approximation of reaction
temperature.
Table 4.2. Synthetic conditions for RbBaxSr2-xNb3O10 compounds using dry microwave synthesis.
Theor. Ba Experim. Mass (g) Mass (g) Mass (g) Mass (g) Rb : Ba : Sr : Nb Heating Time
content Ba Content Ba(NO3)2 SrCO3 Nb2O5 Rb2CO3 Molar Ratio and Power %
8 min at 100%,
0 2 0 0.2223 0.3018 0.1761 1.01 : 0 : 2.00 : 3.02
12 min at 50%
8 min at 100%,
0.1 0.15 0.0196 0.2124 0.302 0.1802 1.03 : 0.09 : 1.91 : 3.02
12 min at 50%
8 min at 100%,
0.25 0.3 0.0488 0.1917 0.2998 0.174 1.00 : 0.25 : 1.73 : 3.00
12 min at 50%
8 min at 100%,
0.5 0.4 0.0991 0.1677 0.3006 0.1744 1.00 : 0.50 : 1.51 : 3.00
12 min at 50%
8 min at 100%,
0.75 0.6 0.1459 0.1399 0.2983 0.1771 1.02 : 0.74 : 1.26 : 2.98
12 min at 50%

4.3 XRD Results and Discussion

Figures 4.1-4.10 and 4.11-15 show XRD patterns for RbCaxSr2-xNb3O10 and

RbSrxBa2-xNb3O10, respectively. Le Bail fits were performed in the TOPAS program using

the Pnma space group and fit parameters are shown in Table 4.3 and 4.4. Lists of indexed

hkl reflections are provided in Appendix A. Increasing Sr content in the Ca/Sr series and

increasing Ba content in the Sr/Ba series results in expansion of the crystal lattice due to

increasing cation size. This is most clearly observed in Figures 4.16 and 4.17 showing

changes in unit cell volume and stacking distance (designated as the c-axis in the unit cell

dimensions), respectively, with changing A-site content. All of the patterns in the Ca/Sr

series showed the presence of a single phase, consistent with the work of Geselbracht et

al.31 and indicative of a homogeneous distribution of cations at the A-site. It should be

noted that the progression in unit cell dimensions in this work does not follow as uniform

a trend as that seen in Geselbracht’s work. Samples in the Ba/Sr series up to a Ba content

of 0.6 formula units also showed the presence of a single phase consistent with the three-

layer structure. Ba content above 0.6 formula units, however, consistently exhibited

61
reflections in the XRD patterns that could not be accounted for using only the three-layer

structure. It is believed the larger Ba2+ ionic radius inhibits formation of a three-layer

structure without substantial Sr content acting to stabilize the structure. CsBa2Nb3O10,

however, can be synthesized successfully owing to the larger Cs+ ionic radius compared to

Rb+. The unaccounted for reflections in the XRD patterns can be fit using a unit cell

consistent with formation of a new phase BaNb2O6.156

62
Table 4.3. Lattice parameters for RbCaxSr2-xNb3O10 compounds obtained from Le Bail fitting using
TOPAS software.
Ca Content Space Group a(Å) b(Å) c(Å) Unit Cell
Volume (Å3)
0* Pnma 7.763(1) 7.860(1) 30.267(3) 1847
0.4* Pnma 7.852(1) 7.744(1) 29.822(7) 1813
0.7* Pnma 7.826(1) 7.724(1) 29.986(4) 1813
0.9* Pnma 7.713(2) 7.787(1) 30.166(4) 1812
1.3* Pnma 7.720(3) 7.779(2) 29.95(1) 1799
1.4* Pnma 7.790(1) 7.711(2) 29.81(1) 1791
1.9* Pnma 7.712(1) 7.730(1) 29.78(1) 1775
2* Pnma 7.690(2) 7.717(3) 29.832(3) 1770
2** Pmna 7.722(1) 7.762(1) 29.768(4) 1784
0** Pmna 7.765(1) 7.781(1) 30.419(5) 1838
0.5** Pmna 7.843(2) 7.771(1) 30.190(3) 1840
1.1** Pmna 7.813(3) 7.771(2) 30.094(8) 1827
1.7** Pmna 7.783(1) 7.746(1) 29.685(4) 1790
* Samples synthesized using the microwave heating method
**Samples synthesized using the molten salt method

Table 4.4. Lattice constants for RbBaxSr2-xNb3O10 compounds.


X Space Group a (Å) b (Å) c (Å)
0 Pnma 7.763(1) 7.860(1) 30.267(3)
0.15 Pnma 7.784(1) 7.863(2) 30.274(9)
0.3 Pnma 7.882(4) 7.770(4) 30.292(8)
0.5 Pnma 7.85(1) 7.80(1) 30.27(2)
0.6 Pnma 7.843(3) 7.775(3) 30.557(9)

RbCa2Nb3O10

4 8 12 16 20 24 28 32 36 40 44 48 52 56 60
2θ (Degrees)

Figure 4.1. XRD pattern for RbCa2Nb3O10. Le Bail fit parameters are shown in Table 4.2.

63
RbCa1.9Sr0.1Nb3O10

4 8 12 16 20 24 28 32 36 40 44 48 52 56 60
2θ (Degrees)

Figure 4.2. XRD pattern for RbCa1.9Sr0.1Nb3O10. Le Bail fit parameters are shown in Table 4.2.

RbCa1.5Sr0.5Nb3O10

4 8 12 16 20 24 28 32 36 40 44 48 52 56 60
2θ (Degrees)

Figure 4.3. XRD pattern for RbCa1.5Sr0.5Nb3O10. Le Bail fit parameters are shown in Table 4.2.

64
RbCa1.4Sr0.6Nb3O10 MSS

4 8 12 16 20 24 28 32 36 40 44 48 52 56 60
2θ (Degrees)

Figure 4.4. XRD pattern for RbCa1.4Sr0.6Nb3O10. Le Bail fit parameters are shown in Table 4.2.

RbCa1.3Sr0.7Nb3O10

4 8 12 16 20 24 28 32 36 40 44 48 52 56 60
2θ (Degrees)

Figure 4.5. XRD pattern for RbCa1.3Sr0.7Nb3O10. Le Bail fit parameters are shown in Table 4.2.

65
RbCa1.1Sr0.9Nb3O10

4 8 12 16 20 24 28 32 36 40 44 48 52 56 60
2θ (Degrees)

Figure 4.6. XRD pattern for RbCa1.1Sr0.9Nb3O10. Le Bail fit parameters are shown in Table X.

RbCa0.9Sr1.1Nb3O10 MSS

4 8 12 16 20 24 28 32 36 40 44 48 52 56 60
2θ (Degrees)

Figure 4.7. XRD pattern for RbCa0.9Sr1.1Nb3O10. Le Bail fit parameters are shown in Table 4.2.

66
RbCa0.7Sr1.3Nb3O10

4 8 12 16 20 24 28 32 36 40 44 48 52 56 60
2θ (Degrees)

Figure 4.8. XRD pattern for RbCa0.7Sr1.3Nb3O10. Le Bail fit parameters are shown in Table 4.2.

RbCa0.4Sr1.6Nb3O10

4 8 12 16 20 24 28 32 36 40 44 48 52 56 60
2θ (Degrees)

Figure 4.9. XRD pattern for RbCa0.4Sr1.6Nb3O10. Le Bail fit parameters are shown in Table 4.2.

67
RbSr2Nb3O10

4 8 12 16 20 24 28 32 36 40 44 48 52 56 60
2θ (Degrees)

Figure 4.10. XRD pattern for RbSr2Nb3O10. Le Bail fit parameters are shown in Table 4.2.

RbBa0.15Sr1.85Nb3O10

4 8 12 16 20 24 28 32 36 40 44 48 52 56 60
2θ (Degrees)

Figure 4.11. XRD pattern for RbBa0.15Sr1.85Nb3O10. Le Bail fit parameters are shown in Table 4.3.

68
RbBa0.3Sr1.7Nb3O10

4 8 12 16 20 24 28 32 36 40 44 48 52 56 60
2θ (Degrees)

Figure 4.12. XRD pattern for RbBa0.3Sr1.7Nb3O10. Le Bail fit parameters are shown in Table 4.3.

RbBa0.4Sr1.6Nb3O10

4 8 12 16 20 24 28 32 36 40 44 48 52 56 60
2θ (Degrees)

Figure 4.13. XRD pattern for RbBa0.4Sr1.6Nb3O10. Le Bail fit parameters are shown in Table 4.3.

69
RbBa0.5Sr1.5Nb3O10

4 8 12 16 20 24 28 32 36 40 44 48 52 56 60
Figure 4.14. XRD pattern for RbBa0.5Sr1.5Nb3O10. Le Bail fit parameters are shown in Table 4.3.

RbBa0.6Sr1.4Nb3O10

4 8 12 16 20 24 28 32 36 40 44 48 52 56 60
2θ (Degrees)

Figure 4.15. XRD pattern for RbBa0.6Sr1.4Nb3O10. Le Bail fit parameters are shown in Table 4.3.

70
1860
Unit Cell Volume vs. Ca content
1850
Unit Cell Volume ,(Å^3)

1840
Microwave Samples
1830
Molten Salt Samples
1820
1810
1800
1790
1780
1770
1760
1 0 1.5 0.5 2 2.5
Ca content, formula units
Figure 4.16. Graph of Unit Cell Volume vs. Ca content in RbCa xSr2-xNb3O10 compounds calculated
using lattice parameters from Le Bail fits. Note the gradual decrease in unit cell volume with
increasing Ca content due to the smaller ionic radius of Ca 2+ compared to Sr2+.

Stacking Distance vs. Ca content


30.3
30.2 Microwave Samples
c(Å) first stacking peak

Molten Salt Samples


30.1
30
29.9
29.8
29.7
29.6
29.5
29.4
0 0.5 1 1.5 2 2.5
Experimental Ca content

Figure 4.17. Graph of stacking distance in RbCa xSr2-xNb3O10 compounds vs. Ca content. The stacking
distance was obtained using the position of the (002) reflection. Note the gradual decrease in stacking
distance with increasing Ca content due to the smaller ionic radius of Ca 2+ vs. Sr2+.

71
4.4 EDS and SEM Results and Discussion

EDS data for the Ca, Ba, and Nb contents shown in Tables 4.5 and 4.6 were used

to determine the composition of each sample in the Ca/Sr and Ba/Sr series. The

composition was calculated using average ratios of Ca/Ba : Nb since Nb content within the

three-layer structure is constant across all samples. The measured Sr content was not used

due to overlap of Rb and Sr in EDS spectra (see Figure 4.18) leading to highly distorted

values. Errors for the calculated composition were determined using Fieller’s method157

of calculating confidence intervals for the ratio of two means. The experimental sample

compositions were in close alignment with theoretical compositions, though

experimentally calculated Ca content was consistently slightly higher than theoretical Ca

content. It is believed this is the result of an inherent error in atomic compositional

measurement on the instrument as evidenced by the experimental value of 2.1 for the pure

RbCa2Nb3O10 sample. This inherent imprecision was considered in subsequent analysis.

Table 4.5. Atom percent results for RbCaxSr2-xNb3O10 obtained using EDS. Ca content was estimated
using ratio of Nb:Ca due to clean separation of peaks in EDS spectra. Ca content error was calculated
using Fieller’s method of estimated confidence intervals between ratios.
Intended Ca Rb (%) error Ca (%) error Sr (%) error Nb (%) error Ca units error
2 7 1 5.6 0.4 0 0 8 1 2.1 0.3
1.75 6 1 7.4 0.4 2 1 12 2 1.9 0.3
1.5 5 0.9 5.6 0.3 3 1 11 2 1.5 0.3
1.25 7 1 4.9 0.3 4 2 11 2 1.3 0.3
1 4.7 0.8 2.1 0.2 6 2 10 2 0.7 0.2
0.75 5.1 0.9 2.9 0.2 7 2 12 2 0.7 0.2
0.5 4.3 0.8 1.7 0.2 5 2 10 2 0.9 0.2
0.25 2.2 0.9 0.6 0.1 5 2 5 1 0.4 0.1
0* 3.8 0.8 0 0 9 2 10 2 NA NA
*Performing the same analysis using Sr content in place of Ca content leads to a Sr unit value of 2.7
(compared to theoretically 2.0) illustrating the distorting effect from the overlapping Rb signal in the EDS
spectra.

72
Table 4.6. Atom percent results for RbBaxSr2-xNb3O10 obtained using EDS. Ba content was estimated
using ratio of Nb:Ca due to clean separation of peaks in EDS spectra. Ba content error was calculated
using Fieller’s method of estimated confidence intervals between ratios.
Intended Ba Content % Ba error % Nb error % Sr error Ba Units error
0 0 0 8 1 6 1 0 0
0.1 0.2 0.1 4 1 3 1 0.15 0.08
0.25 0.8 0.2 7 1 5 2 0.3 0.1
0.4 0.5 0.2 4 1 3 1 0.4 0.1
0.5 1.8 0.3 11 2 6 2 0.5 0.1
0.75 1.2 0.3 6 1 3 1 0.6 0.2

cps/eV
RbCa1Sr1Nb3O10 1

1.8

1.6

1.4

1.2

C Sr
1.0 Nb O Rb Nb Ca Rb Sr Nb

Ca

0.8

0.6

0.4

0.2

0.0
2 4 6 8 10 12 14
keV

Figure 4.18. EDS spectrum for RbCaSrNb3O10. Note the overlapping Rb and Sr signals.

SEM images in Figure 4.19 clearly show that the triple layer compounds form thin

platelets with varying lateral sizes up to several micrometers. This is consistent with

literature reports on the morphology of three-layered perovskites.158-159 Comparisons of

samples synthesized using dry microwave heating and those using molten salts show dry

73
samples exhibit particle sizes substantially larger than molten salt samples. Dry samples

produced particles up to ~5 μm while molten salt samples produced particle sizes up to ~2-

3 μm. This is believed to be the result of increased particle diffusion within molten salts

which allow for reactants to come into contact with greater ease thus forming the desired

layered niobate, but concurrently inhibiting formation of large particles. SEM images of

attempted RbBaxSr2-xNb3O10 syntheses (Figure 4.20) show the presence of expected

platelets, but also rod-like impurities thought to be from formation of BaNb2O6 seen in

XRD patterns and consistent with SEM images for BaNb2O6 published previously.160

These impure samples were not used in subsequent analysis.

a) b)

c) d)
Figure 4.19. SEM images of RbCa2Nb3O10 synthesized under dry microwave (a) and molten salt (b)
conditions and RbSr2Nb3O10 synthesized under dry microwave (c) and molten salt (d) conditions. Note
the increased platelet size seen in dry microwave samples compared to molten salt samples.

74
Figure 4.20. SEM images for attempted RbBa 1.75Sr0.25Nb3O10 sample. Note the combination of
expected platelets and the presence of rod-like impurities in the left image resulting from formation of
the BaNb2O6 phase. The higher magnification image on the right shows the rod-like nature of the
impurity particles.

4.5 ATR-IR Results and Discussion

IR spectra gathered for the mixed A-site samples exhibit two absorption bands

characteristic of the perovskite structure. The strongest is a sharp peak in the 910-920 cm-
1
range due to the asymmetric stretching mode of the terminal Nb-O bond with double bond

character while the second peak occurs in the 690-770 cm-1 range and is associated with

the interior NbO6 octahedral Nb—O—Nb asymmetric stretching model.161-162 The

position of the interior NbO6 peak was found to be highly dependent on composition with

RbCa2Nb3O10 having a peak at 768.9 cm-1, RbSr2Nb3O10 at 738.4 cm-1, and

RbBa0.6Sr1.4Nb3O10 at 691.2 cm-1 while the peak at 910-920 cm-1 shows no significant shift.

Mixed cation samples show a near linear progression in peak position with increasing

cation size seen in both the Ca/Sr and Sr/Ba series of samples as seen in Figure 4.23.

Raman spectral analysis of various niobium oxides (including D-J KCa2Nb3O10 and

HCa2Nb3O10) from Jehng et al.153 clearly illustrated the correlation in stretching mode
75
wavenumber to NbO6 octahedral distortion with bands in the 600-800 cm-1 corresponding

to lightly distorted octahedra and bands in the 850-1000 cm-1 range corresponding to highly

distorted octahedra. Therefore, the IR results here indicate A-site modification does not

significantly alter the distortion of the exterior NbO6 octahedra, but increasing cation size

in the progression from Ca2+ to Sr2+ to Ba2+ leads to reduced octahedral distortions in the

interior NbO6 octahedra. This is further confirmed by the 93Nb NMR results to be discussed

in section 4.6.

RbCaxSr2-xNb3O10 IR vs. Ca Content


775
Wavenumber, cm-1

765

755

745 y = 14.586x + 741.25

735
0 0.5 1 1.5 2 2.5

Ca Content, formula units

RbBaxSr2-xNb3O10 IR vs Ba Content
750
Wavenumber, cm-1

740
730 y = -80.921x + 745.65
720
710
700
690
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Ba Content, formula units

Figure 4.23. Graphs of IR peak position vs. Ca content (Top) and Ba content (Bottom).

76
4.6 NMR Results and Discussion

The results presented up to this point all indicate the distribution of cations in mixed

Ca/Sr and Ba/Sr compounds to be homogeneous with gradual progressions seen in XRD

and IR parameters. These characterization techniques, however, are lacking with respect

to characterizing disorder on the local scale. NMR is unique in its ability to characterize

local environments and, as will be shown, indicates an interesting dichotomy in which

despite evidence of global homogeneity, cation distribution on the local scale exhibits clear
93
heterogeneity. Figures 4.24-4.34 and 4.35-4.38 show the Nb WURST-QCPMG static

spectra (at both 9.4 and 14.1 T) and MQMAS spectra, respectively, for samples in the Ca/Sr

and Ba/Sr series. Static NMR data were used as the primary approach for analyzing the

different 93Nb environments when such environments could be easily distinguished in the

spectra. In cases where the spectral resolution needed for distinguishing multiple

environments was not present, 3QMAS was employed as a supplementary technique.

77
Figure 4.24. 93Nb Static WURST spectra for RbSr2Nb3O10 at 9.4 T (Top) and 14.1 T (Bottom). Blue -
Experimental Spectrum; Black - Total Fit; Green - Exterior Site ;and Red - Interior Site.

78
Figure 4.25. 93Nb Static WURST spectra for RbCa0.4Sr1.6Nb3O10 at 9.4 T (Top) and 14.1 T (Bottom).
Blue - Experimental Spectrum; Black - Total Fit; Green - Exterior Site 1 (Sr-like); Purple - Exterior
Site 2 (Ca-like); and Red - Interior Site.

79
Figure 4.26. 93Nb Static WURST spectra for RbCa0.5Sr1.5Nb3O10 at 9.4 T (Top) and 14.1 T (Bottom).
Blue - Experimental Spectrum; Black - Total Fit; Green - Exterior Site 1 (Sr-like); Purple - Exterior
Site 2 (Ca-like); Yellow – Impurity; and Red - Interior Site.

80
Figure 4.27. 93Nb Static WURST spectra for RbCa0.7Sr1.3Nb3O10 at 9.4 T (Top) and 14.1 T (Bottom).
Blue - Experimental Spectrum; Black - Total Fit; Green - Exterior Site 1 (Sr-like); Purple - Exterior
Site 2 (Ca-like); and Red - Interior Site.

81
Figure 4.28. 93Nb Static WURST spectra for RbCa0.9Sr1.1Nb3O10 at 9.4 T (Top) and 14.1 T (Bottom).
Blue - Experimental Spectrum; Black - Total Fit; Green - Exterior Site 1 (Sr-like); Purple - Exterior
Site 2 (Ca-like); and Red - Interior Site.

82
Figure 4.29. 93Nb Static WURST spectra for RbCa1.1Sr0.9Nb3O10 at 9.4 T (Top) and 14.1 T (Bottom).
Blue - Experimental Spectrum; Black - Total Fit; Green - Exterior Site 1 (Sr-like); Purple - Exterior
Site 2 (Ca-like); and Red - Interior Site.

83
Figure 4.30. 93Nb Static WURST spectra for RbCa1.4Sr0.6Nb3O10 at 9.4 T (Top) and 14.1 T (Bottom).
Blue - Experimental Spectrum; Black - Total Fit; Green - Exterior Site 1 (Sr-like); Purple - Exterior
Site 2 (Ca-like); and Red - Interior Site.

84
Figure 4.31. 93Nb Static WURST spectra for RbCa1.5Sr0.5Nb3O10 at 9.4 T (Top) and 14.1 T (Bottom).
Blue - Experimental Spectrum; Black - Total Fit; Green - Exterior Site 1 (Sr-like); Purple - Exterior
Site 2 (Ca-like); and Red - Interior Site.

85
Figure 4.32. 93Nb Static WURST spectra for RbCa1.7Sr0.3Nb3O10 at 9.4 T (Top) and 14.1 T (Bottom).
Blue - Experimental Spectrum; Black - Total Fit; Green - Exterior Site 1 (Sr-like); Purple - Exterior
Site 2 (Ca-like); and Red - Interior Site.

86
Figure 4.33. 93Nb Static WURST spectra for RbCa1.9Sr0.1Nb3O10 at 9.4 T (left) and 14.1 T (right). Blue
- Experimental Spectrum; Black - Total Fit; Green - Exterior Site 1 (Sr-like); Purple - Exterior Site
2 (Ca-like); and Red - Interior Site.

87
Figure 4.34. 93Nb Static WURST spectra for RbCa2Nb3O10 at 9.4 T (Top) and 14.1 T (Bottom). Blue -
Experimental Spectrum; Black - Total Fit; Purple - Exterior Site and Red - Interior Site.

The double shearing transformation used in our MQMAS processing allows for

the separation of signals based on CQ values and was thus used to confirm the presence of

multiple unique environments associated with exterior Nb octahedra. First moment

analysis was employed to extract CQ values for several samples, as shown in Table 4.7.

These values were consistent with those obtained from static fits of spectra with better
88
resolved features. Another useful feature of the obtained MQMAS spectra is the

consistent slope and width in the 93Nb signals. The observed slope is indicative of a

distribution of CQ, consistent with both the disordered nature of these materials and the

static NMR spectra. Using this combination of NMR experiments, fits were obtained

leading to unique observations regarding the structure of mixed cation perovskites.

MQMAS lineshape analysis was employed to determine CQ and η, then subsequently

used to extract appropriate static 93Nb fit parameters in Table 4.8.

Table 4.7. First moment analysis for RbCa xSr2-xNb3O10 compounds. The obtained CQ and η values
were calculated using Equation 2.42-45.
Sample CQ, MHz η <v1>, ppm

RbCa0.9Sr1.1Nb3O10 29.2 0.32 -197

RbCa0.7Sr1.3Nb3O10 30.5 0.34 -215

44.4 0.06 -382

RbCa0.5Sr1.5Nb3O10 28 0.24 -181

RbCa0.4Sr1.6Nb3O10 35.9 0.21 -297

44.9 0.01 -388

89
Figure 4.35. Contour plot of 93Nb 3QMAS 2D spectrum of RbCa0.4Sr1.6Nb3O10 with double shearing
scheme. The horizontal axis corresponds to the isotropic dimension and vertical axis to the anisotropic
dimension. Note the observed signal positions along F2 are not equal to the isotropic chemical shifts
in static spectra, as illustrated in Equation 2.42.

Figure 4.36. Contour plot of 93Nb 3QMAS 2D spectrum of RbCa0.5Sr1.5Nb3O10 with double shearing
scheme. The horizontal axis corresponds to the isotropic dimension and vertical axis to the anisotropic
dimension. Note the observed signal positions along F2 are not equal to the isotropic chemical shifts in
static spectra, as illustrated in Equation 2.42.
90
Figure 4.37. Contour plot of 93Nb 3QMAS 2D spectrum of RbCa0.7Sr1.3Nb3O10 with double shearing
scheme. The horizontal axis corresponds to the isotropic dimension and vertical axis to the anisotropic
dimension. Note the observed signal positions along F2 are not equal to the isotropic chemical shifts in
static spectra, as illustrated in Equation 2.42.

Figure 4.38. Contour plot of 93Nb 3QMAS 2D spectrum of RbCa0.9Sr1.1Nb3O10 with double shearing
scheme. The horizontal axis corresponds to the isotropic dimension and vertical axis to the anisotropic
dimension. Note the observed signal positions along F2 are not equal to the isotropic chemical shifts in
static spectra, as illustrated in Equation 2.42.

91
Assuming the homogeneous cation distribution observed in long-range and bulk

properties, it would be expected that the 93Nb spectra would contain one signal from the

exterior NbO6 octahedra and a second from the interior NbO6 octahedra with a gradual

change in CQ values as composition changes. However, as mentioned previously, this is

not the case. Rather, spectra consistently exhibit multiple signals associated with exterior

octahedra while there is one identifiable signal associated with interior octahedra. It should

be noted that the observation of a single interior 93Nb signal does not negate the absence

of a second. The substantially larger CQ value in the interior site results in signal intensity

being spread over a much larger range such that low population signals may not be readily

distinguished. Another consequence of the low intensity signal arising from the interior

environments is a lack of signal to noise necessary for more in-depth analysis such as that

which will be presented for the exterior environments.

92
Table 4.9. RbCaxSr2-xNb3O10 NMR Fit parameters. Errors in the last digit are shown in parentheses and were calculated for population and CQ
values using chi-square minimization.
X Site Pop. (%) δiso (ppm) CQ (MHz) η Ω (ppm) κ α (Deg) β (Deg) γ (Deg)
0 Ext 1 62.8(3) -963 45.1(4) 0.04 831 0.83 0 0 0
Int 37.2(4) -1043 93(1) 0.01 490 0.97 0 0 0
0.4 Ext 1 27(1) -953 35.9(8) 0.21 1091 0.65 44.1 9 24.9
Ext 2 41(1) -955 44.9(7) 0.01 851 0.9 0 0 0
Int 33(2) -1051 92(2) 0.04 875 0.53 38.7 6.3 33.9
0.5 Ext 1 33(3) -931 32(1) 0.27 1118 0.59 44.8 1.9 7.9
Ext 2 4.2(7) -954 28(2) 0.24 1105 0.7 92.2 12.9 16.3
Ext 3 35(1) -955 44.5(6) 0.07 852 0.91 0 0 0
Int 27.2(9) -1068 93(3) 0.06 984 0.27 -14.6 15.9 62.1
0.7 Ext 1 24.5(8) -976 30.5(7) 0.34 1114 0.6 100 7.2 87.4
Ext 2 50.5(5) -950 44.4(5) 0.06 849 0.92 0 0 0
Int 25(4) -1068 92(2) 0.06 978 0.29 72.8 15.9 61
0.9 Ext 1 21(3) -981 29.2(4) 0.32 956 0.51 44.6 3.64 19.1
Ext 2 45(2) -953 45(1) 0.06 817 0.91 31.7 0.73 45.3
93

Int 33(2) -1069 93(2) 0.06 544 0.87 0 0 0


1.1 Ext 1 60(1) -954 33(1) 0.27 978 0.51 -1.7 3.3 15.2
Ext 2 18.6(7) -958 44(1) 0.12 861 0.87 51 5.2 -8
Int 21(5) -1071 105(3) 0.18 683 0.73 42.4 17 41.1
1.4 Ext 1 40(2) -937 32.6(7) 0.29 931 0.51 20.9 2 26.4
Ext 2 38(2) -905 44(1) 0.18 870 0.82 10.3 0.3 16
Int 21(7) -1071 104(4) 0.19 757 0.61 14.3 14.7 48.8
1.5 Ext 1 49(1) -962 33.7(5) 0.26 872 0.63 0 2 26.4
Ext 2 29.2(9) -953 43.5(6) 0.18 765 0.81 10.3 0.3 16.1
Int 21(6) -1034 111(2) 0.16 816 0.4 13.9 13.9 54.3
1.7 Ext 1 53(2) -974 31(1) 0.28 904 0.67 33.1 2.2 44.5
Ext 2 15(4) -926 44.5(4) 0.06 842 0.92 12.2 0.2 10
Int 33(3) -940 111(5) 0.17 905 0.1 1.1 16.4 43.5
1.9 Ext 1 64(1) -981 30.1(9) 0.21 892 0.74 0.9 1 10.6
Int 35(2) -992 110(2) 0.17 795 0.66 14.5 7.1 44.1
2 Ext 1 59(1) -987 28.8(7) 0.19 910 0.64 45 4.6 27
Int 40(3) -1102 109(4) 0.11 1060 0.86 0 0 0
For easier discussion and analysis of the numerous parameters for the mixed Ca/Sr

series, Figures 4.39-4.43 show graphs of Ca content vs CQ, Population, Ω, η, and κ values,

respectively.

93Nb CQ vs. Ca Content


50
125
45
Exterior CQ, MHz

Interior CQ, MHz


115
40
105
35 95

30 85

25 75
0 0.5 1 1.5 2
Ca Content, formula units

Exterior 1 Exterior 2 Interior

Figure 4.39. Graph of 93Nb CQ (in MHz) vs Ca content for Exterior Site 1 (blue), Exterior Site 2
(orange), and Interior Site (grey).

94
Site Population vs. Ca Content
70

Population, % 60
50
40
30
20
10
0
0 0.5 1 1.5 2
Ca Content, formula units

Exterior 1 Exterior 2 Interior

Figure 4.40. Graph of 93Nb Population vs Ca content for Exterior Site 1 (blue), Exterior Site 2 (orange),
and Interior Site (grey).

CSA Span vs. Ca Content


1200
CSA Span (Ω), ppm

1000

800

600

400

200

0
0 0.5 1 1.5 2
Ca Content, formula units

Exterior 1 Exterior 2 Interior

Figure 4.41. Graph of 93Nb CSA span (Ω, ppm) vs Ca content for Exterior Site 1 (blue), Exterior Site
2 (orange), and Interior Site (grey).

95
Asymmetry Parameter vs. Ca Content
0.4
0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
0 0.5 1 1.5 2
Ca Content, formula units

Exterior 1 Exterior 2 Interior

Figure 4.42. Graph of 93Nb asymmetry parameter (η) value vs Ca content for Exterior Site 1 (blue),
Exterior Site 2 (orange), and Interior Site (grey).

Skew vs. Ca Content


1.2

1
Skew (κ) value

0.8

0.6

0.4

0.2

0
0 0.5 1 1.5 2
Ca Content, formula units

Exterior 1 Exterior 2 Interior

Figure 4.43. Graph of 93Nb skew (κ) value vs Ca content for Exterior Site 1 (blue), Exterior Site 2
(orange), and Interior Site (grey).

96
Fits of exterior site signals in the Ca/Sr series consistently show one environment

with a CQ value close to that of the pure Ca form (~28 MHz) and a second environment

with a CQ value close to that of pure Sr form (~45 MHz). The Ca-like environment does

see significant fluctuations in CQ value going as high as 35.9 MHz in the Ca0.4 sample and

consistently being larger than that seen in the pure Ca form. The Sr-like environment,

however, sees less than 5% fluctuation in CQ values up to Ca1.9 (at which point no Sr-like

environment can be distinguished). Fits of the interior 93Nb signals, while being assigned

to single 93Nb environments, still do not exhibit the same gradual changes observed in long-

range measurements, but instead can be treated as either a Ca-like or Sr-like environment.

Interestingly, despite clear evidence of heterogenous distribution of A-site cations,

populations of the observed environments do not correlate to the composition determined

from EDS data suggesting that the A-site cation effects the EFG tensor of octahedra beyond

those directly adjacent to that particular A-site. This is illustrated in Figure 4.40 showing

a comparison of site populations for varying Ca contents. Looking at the exterior sites,

while there is a general trend where the population of Ca-like environments increases with

increasing Ca content, individual sample analysis does not correspond to composition

exactly. For instance, RbCa0.5Sr1.5Nb3O10 would theoretically have two exterior sites with

a population ratio of 1:3 for Ca-like:Sr-like, but experimentally this ratio is approximately

1:1.

The degree of octahedral tilting in both the xy plane and along the z-axis for the

interior NbO6 octahedra is correlated to CQ values with larger tilting leading to larger EFG

97
values. Thus Ca-like environments correspond to more highly distorted octahedra than Sr-

like environments. Considering the smaller ionic radius of Ca2+, tilting of octahedra is

expected in order for Ca2+ to coordinate stably to interior octahedral O2-. EFG values in

the exterior site are instead the result of out of center displacement of the Nb atom in the

octahedra. Considering the double bond character of the terminal Nb=O bond, this is to be

expected. As evidenced by the lack of shifting in the terminal Nb=O IR absorption band,

A-site does not lead to significant changes in Nb=O bond length. Changing EFG values

are then attributed to shifting of the equatorial oxygen atom position and thus less out center

displacement of the Nb atom with decreasing A-site cation size. Spectra collected for

samples in the Ba/Sr series were analyzed in a similar fashion to the Ca/Sr series.

98
Figure 4.44. 93Nb static WURST spectra for RbBa0.15Sr1.85Nb3O10 at 9.4 T (Top) and 14.1 T (Bottom).
See Table 4.8 for detailed fit parameters. Experimental spectrum is shown in blue and overall fit is
shown in red.

99
Figure 4.45. 93Nb static WURST spectra for RbBa0.3Sr1.7Nb3O10 at 9.4 T (Top) and 14.1 T (Bottom).
See Table 4.8 for detailed fit parameters. Experimental spectrum is shown in blue and overall fit is
shown in red.

100
Figure 4.46. 93Nb static WURST spectra for RbBa0.4Sr1.6Nb3O10 at 9.4 T (Top) and 14.1 T (Bottom).
See Table 4.8 for detailed fit parameters. Experimental spectrum is shown in blue and overall fit is
shown in red.

101
Figure 4.47. 93Nb static WURST spectra for RbBa0.6Sr1.4Nb3O10 at 9.4 T (Top) and 14.1 T (Bottom).
See Table 4.8 for detailed fit parameters. Experimental spectrum is shown in blue and overall fit is
shown in red.

102
Table 4.10. RbBaxSr2-xNb3O10 93Nb Static NMR Fit parameters. Errors in the last digit are shown in parentheses and were calculated for
population and CQ values using chi-square minimization.
X Site Pop. (%) δiso (ppm) CQ (MHz) η Ω (ppm) κ α (Deg) β (Deg) γ (Deg)

0 Ext 1 62.8(3) -963 45.1(4) 0.04 831 0.83 0 0 0


Int 37.2(4) -1043 93(1) 0.01 490 0.97 0 0 0
0.15 Ext 1 61(2) -959 45.7(9) 0.08 800 0.86 0 0 0
Ext 2 8.8(6) -970 53(1) 0.14 900 0.92 0 0 0
Int 1 22(3) -1045 92(3) 0.05 586 0.46 0 0 0
Int 2 5.4(9) -1051 90(2) 0.02 350 0.79 0 0 0
Impurity 2(1) -945 21(1) 0.8 110 0.08 0 0 0
0.3 Ext 1 50(3) -960 45(2) 0.14 795 0.88 0 0 0
Ext 2 17(1) -965 55(2) 0.17 905 0.92 0 0 0
Int 1 16(1) -1056 93(3) 0.06 560 0.54 0 0 0
Int 2 10(1) -1068 87(2) 0.05 410 0.70 0 0 0
Impurity 6.9(7) -969 22(1) 0.83 110 0.08 0 0 0
0.4 Ext 1 49(4) -958 45.3(8) 0.19 805 0.90 0 0 0
103

Ext 2 18(2) -950 53.5(9) 0.11 890 0.90 0 0 0


Int 1 16(3) -1060 93(4) 0.06 600 0.62 0 0 0
Int 2 11(1) -1066 88(3) 0.03 445 0.72 0 0 0
Impurity 5.3(5) -980 20(1) 0.82 110 0.16 0 0 0
0.6 Ext 1 39(2) -960 43(1) 0.21 810 0.86 0 0 0
Ext 2 31(3) -945 55(1) 0.2 880 0.80 0 0 0
Int 1 5.8(6) -1075 93(3) 0.07 560 0.84 0 0 0
Int 2 19(1) -1039 85(2) 0.09 500 0.56 0 0 0
Impurity 5.8(7) -963 21.4(8) 0.85 120 0.13 0 0 0
As seen in Figures 4.44-4.47 and Table 4.8, RbBaxSr2-xNb3O10 93Nb spectra and the

corresponding fit spectra indicate a clear heterogeneity between Sr-like environments and

Ba-like environments. This is further illustrated in graphs of CQ, population, CSA, skew,

and asymmetry parameter vs. Ba content in Figures 4.48-4.52, respectively. Using work

regarding changes in the 93Nb EFG of CsSr2Nb3O10 and CsBa2Nb3O10, the samples in this

series were fit using five total 93Nb environments.98 The sites designated Ext 1 and Ext 2

in Table 4.8 correspond to Sr-like and Ba-like environments, respectively, with CQ values

typically 44-46 MHz in Ext 1 and 53-55 MHz in Ext 2. As discussed in the Ca/Sr analysis,

this increase in CQ corresponds to decreasing degrees of octahedral tilting and Nb

displacement in the exterior octahedra. Unlike the exterior octahedra, decreasing degrees

of octahedral tilting in interior octahedra results in decreasing CQ values. The sites

designated Int 1 and Int 2 in Table 4.9 correspond to Sr-like and Ba-like environments,

respectively, with CQ values typically 92-94 MHz in Int 1 and 86-88 MHz in Int 2

indicating the Int 2 octahedra have decreased octahedral tilting compared to Int 1. It

should be noted these ranges are larger than those seen in pure Sr or Ba compounds, and is

a result of the disordered nature of cation distribution. The substantially smaller CQ value

of the interior Ba-like octahedra compared to Ca-like octahedra lead to greater signal

intensity and thus the two interior sites can be clearly identified in the Ba/Sr series, but not

in the Ca/Sr series. The final site included in the calculated spectra has been designated an

impurity due to its low CQ (~20 MHz), CSA (~110 ppm), and population (~5%) being

inconsistent with the layered perovskite structure. This impurity could possibly be due to

104
a BaNb2O6 phase. This phase is seen in XRD patterns for attempted syntheses of higher

Ba content samples, and if it is amorphous and/or present in small quantities in low Ba

content samples it would not be observed in XRD, but would be present in the 93Nb spectra.

93Nb CQ vs. Ba Content


100

90

80
CQ, MHz

Ext 1
70
Ext 2
60 Int 1
50 Int 2

40
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Ba Content, formula units

Figure 4.48. Graph of 93Nb CQ vs. Ba content for Exterior Site 1 (blue), Exterior Site 2 (orange),
Interior Site 1 (grey), and Interior Site 2 (yellow).

105
Site Population vs. Ba Content
70
60
50
Population, %

40 Ext 1
30 Ext 2

20 Int 1

10 Int 2

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Ba Content, formula units

Figure 4.49. Graph of 93Nb CQ vs. Ba content for Exterior Site 1 (blue), Exterior Site 2 (orange),
Interior Site 1 (grey), and Interior Site 2 (yellow).

CSA Span vs. Ba Content


1000

900
CSA Span (Ω), ppm

800

700 Ext 1
600 Ext 2
Int 1
500
Int 2
400

300
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Ba Content, formula units

Figure 4.50. Graph of 93Nb CQ vs. Ba content for Exterior Site 1 (blue), Exterior Site 2 (orange),
Interior Site 1 (grey), and Interior Site 2 (yellow).

106
Skew vs. Ba Content
1

0.9
Skew (κ) value

0.8
Ext 1
0.7
Ext 2
0.6 Int 1
0.5 Int 2

0.4
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Ba Content, formula units

Figure 4.51. Graph of 93Nb CQ vs. Ba content for Exterior Site 1 (blue), Exterior Site 2 (orange),
Interior Site 1 (grey), and Interior Site 2 (yellow).

Asymmetry Parameter vs. Ba Content


0.25
Asymmetry Parameter (η) value

0.2

0.15 Ext 1
Ext 2
0.1
Int 1
0.05 Int 2

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Ba Content, formula units

Figure 4.52. Graph of 93Nb CQ vs. Ba content for Exterior Site 1 (blue), Exterior Site 2 (orange),
Interior Site 1 (grey), and Interior Site 2 (yellow).

107
As will be discussed in Chapter 5, heterogeneous cation distribution can often be

attributed to preference for unique crystallographic sites. In the three-layer structure

studied here, though, there is only one unique crystallographic A-site and thus

heterogeneity observed on the local scale cannot be explained based on preference on the

global scale. Considering the rapid nature of these compounds’ syntheses, it is possible

that longer reaction time such as that used in conventional methods would allow for greater

diffusion of cations and thus heterogeneity would not be observed. While this possibility

most certainly deserves further study, the following chapters will show local-scale

heterogeneity to be a potentially huge advantage in the compositional design of layered

perovskites.
93
As mentioned previously, the resolution of Nb signals arising from interior

octahedra is not at a level necessary to easily distinguish the effects of cation distribution.

Therefore, while this analysis is primarily focused on cation distribution in terms of

exterior octahedral effects, the relative magnitude of effects on exterior and interior sites

cannot be accurately compared using current data. Considering Figures 4.39-4.43, clear

heterogeneity is observed most readily in analysis of CQ and η values. Fluctuations in

various fit parameter values may be the result of WURST pulse inhomogeneity, disordered

nature of the compounds, and rapid synthetic methods. Based on the structural

relationships with these parameters discussed in Chapter 2, it is expected they would

exhibit the most readily distinguished effects of heterogeneous distribution as they are

related to both symmetry about the nucleus and internuclear distances. The presence of

108
93
both Sr-like and Ca-like exterior Nb environments suggests there is a heterogeneous

distribution of exterior octahedral distortions, with localized areas of low-distortion (Sr-

like) and other areas with high-distortion (Ca-like). As Chapter 6 will discuss in further

detail, less distorted octahedra lead to weakening of O-H bonds in acid exchanged samples

(i.e. more reactive acid sites). Mixed cation compounds then can exhibit substantially

increased reactivity from these localized regions without altering the majority of the

structure.

4.7 Conclusions

RbSrxCa2-xNb3O10 and RbBaxSr2-xNb3O10 compounds were successfully

synthesized and analyzed using XRD, SEM, EDS, ATR-IR, and NMR. XRD patterns

confirmed formation of the triple layered compounds while yielding lattice parameters

consistent with lattice expansion due to increased cation size. SEM showed the desired

platelet formation with varying uniformity and size depending on the synthetic route while

EDS data were used to calculate the Ca, Sr, or Ba content in each sample. IR spectra

showed an absorption band that shifts linearly with increasing Sr content in the Ca/Sr series

and increasing Ba content in the Ba/Sr series. This band is likely due to changes regarding

the interior Nb octahedra. NMR experiments utilizing the WURST-QCPMG pulse

sequence clearly show the presence of uniquely Ca-like and Sr-like environments in the
93
Ca/Sr series, and uniquely Ba-like and Sr-like environments in the Ba/Sr series. Nb NMR

analysis clearly indicates heterogeneity on the local-scale while bulk properties can be

109
attributed to global-scale homogeneity. As the following chapters will show, local-scale

heterogeneous cation distribution is potentially a huge advantage in the design of new

compounds.

110
Chapter 5: Determination of Cation Site Ordering in Four-
layer HCa2NaNb4-xTaxO13 Compounds Using Solid State 93Nb
and 23Na NMR

5.1 Introduction

Building off the success of studying cation distribution in mixed A-site Dion-

Jacobson niobates, our attention then turned towards the study of cation distribution in

compounds with both mixed A- and B-sites, specifically HCa2NaNb4-xTaxO13. Unlike the

Rb(Ca/Ba)xSr2-xNb3O10 compounds investigated in Chapter 4, there are distinct

crystallographic A- and B-sites in HCa2NaNb4-xTaxO13 that can be accurately studied using

XRD. Unfortunately, due to the rapid synthetic procedure employed and inherent disorder

of these samples, conventional PXRD is insufficient and only a synchrotron source would

provide the signal to noise necessary for Rietveld refinement.163. To further reiterate, NMR

is ideal for disordered systems and was thus employed to determine site-ordering through

observations of 93Nb and 23Na nuclei.

It has been shown that changing the B-site atom can drastically alter the

intercalation behavior and catalytic ability of these materials.31, 55-56 Geselbracht et al.31

studied the acidity of HCa2NbxTa3-xO10 compounds through intercalation and found that

the pyridine intercalation reaction did not work for x=2.5 and 3.0 compounds suggesting a

significant decrease in solid acid strength. For x=0.5-2.0, however, the level of pyridine

intercalation was greater than in the x=0 compound suggesting the ordering of Ta5+ within

the structure has an effect as to whether the solid acid strength is decreased or increased.

111
In work on the water splitting ability of Sr2Ta2O7 and Sr2Nb2O7, Kudo et al.57 found the

water splitting activity and band gaps of the SrTa2O7 and SrNb2O7 compounds to be

significantly different (4.6 eV vs. 3.9 eV, respectively). Subsequent study of

Sr2(Ta1-xNbx)2O7 compounds found photocatalytic activity does not progress linearly with

changing composition.164

While Nb5+ and Ta5+ possess the same charge and approximately equal ionic radii58

(0.64 Ǻ), Nb is more electronegative than Ta (1.6 vs. 1.5 PU, respectively)59 leading to

Nb-O bonds more covalent in character than Ta-O bonds and changes in the energy states

of Ta5+ and Nb5+ compounds.57, 60 Differences in both charge and ionic radius have been

shown to be strong drivers of site ordering in mixed cation compound such as mixed

Nb5+/Ti4+ layered perovskites.61-62 In mixed Nb5+/Ta5+ systems165-166, however, ionic

character has been shown to be the driver of site preference, with the Ta5+ cation occupying

more symmetrical sites due to the more ionic character of Ta-O bonds.

112
Figure 5.1. Tetragonal crystal structure for RbCa2NaNb4O13

In the present study, we have attempted to determine any site preference of Ta in

four layer Rb/HCa2NaNb4-xTaxO13 systems using solid-state NMR techniques. As

illustrated in Figure 5.1, there are two distinct crystallographic B-sites in the four-layer

compound, two interfacial octahedra, Nb(1), and two interior octahedra, Nb(2). X-ray

powder diffraction is unable to give quantitative results as to the populations of Nb5+ and

Ta5+ at each crystallographic site due to the disordered nature of these compounds. NMR,

however, allows for the study of the local structure and is ideal for systems of this sort. To

the best of our knowledge, this is the first reported evidence of site ordering in mixed

Nb5+/Ta5+ D-J perovskites.

Previous reports of the crystal structure of RbCa2NaNb4O1335, 167 suggested Na and

Ca were randomly distributed throughout the perovskite slabs, but to our knowledge no

research has been performed to confirm this. It has been shown that the composition of
113
the alkali cations in layered perovskites has a significant effect on structure and

properties.63 It is logical to assume that distribution of cations in the four-layer systems

would also influence structure and properties. Determining whether the cations are

randomly distributed or if some site preference exists is important information for any

attempt at predicting structure and properties based on composition. To study the

distribution of interstitial cations in the compounds, we performed 23Na NMR experiments

to extract chemical shift and EFG data, and analyzed relative populations to determine

distribution. This new information regarding the distribution of A and B site cations should

help in predicting the behaviors of mixed metal layered perovskites, especially regarding

solid acid behavior.

As this report will show, solid state NMR serves a vital purpose in the determination

of cation ordering within mixed cation compounds. By obtaining static spectra at multiple

fields as well as using Multiple Quantum Magic Angle Spinning (MQMAS) experiments

we could accurately determine the different EFG environments in the compound. Cross-

polarization (CP) and recoupling experiments allowed for the linking of the EFG

environment to the crystallographic position of the Nb octahedra. Using the relative

populations of the static spectra, quantitative determination of ordering within the

compounds was successful and points to a non-random arrangement for both sets of cations

in the mixed Nb/Ta systems.

114
5.2 Experimental
Relevant details of synthetic procedures and characterization techniques are

provided in Chapter 3. Elemental Analysis using ICP-OES was performed by Robertson-

Microlit Laboratories.

5.3 Results and Discussion

5.3.1 XRD, SEM, and Elemental Analysis

XRD patterns for the four-layer Rb samples (Figure 5.2) and protonated samples

(Figure 5.3) confirm formation of the desired structures with the orthorhombic Pnma space

groups, along with a small amount of precursor impurities. Earlier work on the crystal

structure of RbCa2NaNb4O13 determined the structure to be tetragonal35 (P4/mmm space

group, a = 3.8702(1) Ǻ, c = 18.8936(6) Ǻ), however bond valence calculations of this

structure results a global instability index of 0.34. A value above 0.2 indicates an unstable

structure,22 and thus an orthorhombic unit cell is more appropriate than tetragonal. Due to

the similar size of Nb5+ and Ta5+, there is no significant difference in lattice parameters

between the pure Nb5+ and Ta5+ incorporated samples. Both samples’ XRD patterns were

fit using a Pnma space group with lattice parameters doubled compared to the tetragonal

unit cell reported in the literature and lattice parameters a = 7.9162(7) Ǻ, b = 7.7364(6) Ǻ,

and c = 37.729(3) Ǻ for RbCa2NaNb4O13 and a = 7.7328(7) Ǻ b = 7.7178(7) Ǻ, and c =

37.628(3) Ǻ for RbCa2NaNb2.8Ta1.2O13. Ta5+, however, has a stronger x-ray scattering

factor and analysis of the ratio of the reflection intensities for the (100) and (110) hkl

115
reflections indicated successful incorporation of Ta5+ into the crystal lattice. Powder

diffraction, however, was unable to give information on the position of Ta5+ in the

structure. After the acid-exchange was performed, the (00l) stacking reflections were

shifted to a slightly higher 2θ angle because of the decreased size of the interfacial cation.

Detailed Le Bail fit information is given in Table 5.1. It should be noted small stacking

reflections related to the RbCa2Nb3O10 precursor are present in XRD patterns, though the

absence of 93Nb signals associated with this phase in subsequent NMR spectra indicate the

amount of precursor is too low to significantly affect the analysis.

SEM images for the acid exchanged samples (Figure 5.4) show large platelet

formation with lateral dimensions of several micrometers, consistent with literature reports

for other layered perovskites.52 Platelet size does appear to be somewhat smaller in the

Ta5+ containing sample, possibly due to the necessary diffusion of Ta5+ into the crystal

structure acting as a platelet growth inhibitor. The syntheses of Ta containing compounds

in general require a higher reaction temperature than pure Nb compounds suggesting that

a reaction temperature higher than the 1200 oC used may have promoted better platelet

growth. Elemental analysis was performed using ICP-OES from Robertson-Microlit

Laboratories. The weight percentage composition for Nb5+ and Ta5+ in the mixed

Nb5+/Ta5+ sample was 26.26 and 22.98%, respectively. These weight percentages

correspond to an empirical formula of RbCa2NaNb2.8Ta1.2O13, slightly higher than the

theoretical formula of RbCa2NaNb3TaO13 based on the stoichiometry of the precursors

during synthesis.

116
RbCa2NaNb4O13

RbCa2NaNb2.8Ta1.2O13


Figure 5.2. XRD patterns for RbCa2NaNb4O13 (above) and RbCa2NaNb2.8Ta1.2O13 (below). Note that
the pattern for the Ta compound is fit with both the four and three-layer structures represented by
blue and black ticks, respectively.

117
HCa2NaNb4O13

HCa2NaNb2.8Ta1.2O13


Figure 5.3. XRD patterns for HCa2NaNb4O13 (above) and HCa2NaNb2.8Ta1.2O13 (below). Note that the
pattern for the Ta compound is fit with both the four and three-layer structures represented by blue
and black ticks, respectively.

118
Table 5.1. Le Bail method fit lattice parameters, particle size, zero correction, χ2, and Rwp* for XRD
patterns shown in Figures 3 and 4
Sample Space a (Ǻ) b (Ǻ) c (Ǻ) Particle Zero χ2 Rwp
Group Size (Ǻ) Correction (%)
Figure 5.1
RbCa2NaNb4O13 Pnma 7.9162(7) 7.7364(6) 37.729(3) 123.4(4) 0.1202(9) 1.13 16.36
RbCa2Nb3O10 Pnma 7.8710(8) 7.7643(8) 32.119(3) 85.0(4) 0.1202(9) 1.13 16.36
RbCa2NaNb2.8Ta1.2O13 Pnma 7.7328(7) 7.7178(7) 37.628(3) 140.0(8) 0.0096(1) 1.08 20.85
RbCa2Nb3O10 Pnma 7.7935(5) 7.8440(8) 29.845(2) 153(10) 0.0096(1) 1.08 20.85
Figure 5.2
HCa2NaNb4O13 Pnma 7.7259(5) 7.7319(9) 36.818(2) 80.2(1) -0.0193(8) 1.43 15.17
HCa2Nb3O10 Pnma 7.869(1) 7.8353(7) 28.866(2) 87.3(4) -0.0193(8) 1.43 15.17
HCa2NaNb2.8Ta1.2O13 Pnma 7.7296(9) 7.978(1) 36.703(3) 50.0(1) 0.009(1) 3.28 9.37
HCa2Nb3O10 Pnma 7.7853(5) 7.8612(6) 29.124(2) 114.7(6) 0.009(1) 3.28 9.37

Figure 5.4. SEM images of RbCa2NaNb4O13 (top) and RbCa2NaNb2.76Ta1.24O13 (bottom) showing
platelets with lateral dimension of 1-3 μm.

5.3.2 93Nb NMR

Static 93Nb VOCS NMR spectra for HCa2NaNb4O13 and HCa2NaNb2.8Ta1.2O13 at

9.4 T and 14.1 T shown in Figure 5.5 were fit using 4 sites for both the pure Nb5+ and Ta5+

incorporated compounds (detailed site parameters shown in Table 5.3). Three sites had

moderate quadrupolar coupling values (CQ) in the range of 20-30 MHz. These values are

consistent with the interlayer 93Nb environments based on previous work on the three-layer

RbCa2Nb3O10 compound which had an interlayer 93Nb CQ value of 30 MHz. The fourth

site CQ value was substantially larger at approximately 50 MHz. In the three-layer

compound the interior site CQ value is traditionally much larger than the interlayer site (110
119
MHz vs. 30 MHz). Assuming the fourth site is associated with the interior Nb octahedra,

the large difference between the four-layer and three-layer EFG values is consistent with

the different crystal structures of the compounds. In the three-layer compound there is only

one interior Nb(2) octahedra per three-layer slab, whereas in the four-layer compound there

are two interior Nb(2) octahedra. Having two interior octahedra creates a bridging O atom

between the two interior Nb(2) octahedra not seen in the three-layer compound. The axial

interior Nb-O bonds in the three-layer system have been found to be 1.876 Ǻ.168 Assuming

the Nb-O bond lengths presented in previous literature35 for RbCa2NaNb4O13 are not

significantly different from our compounds, the axial interior bridging Nb(2)-O bond

length at 1.99424 Ǻ is approximately 0.15 Ǻ longer than the Nb-O bond in the interior

octahedra of the three-layer compound and close to equatorial Nb-O bond lengths. Based

on the longer Nb-O bond, the CQ value of the interior Nb site would be expected to decrease

substantially due to the site now having a more symmetric environment.

120
HCa2NaNb4O13 14.1 T HCa2NaNb2.8Ta1.2O13
14.1 T

0 -100 -200 -300 -400 0 -100 -200 -300 -400


kHz kHz
a) b)

HCa2NaNb4O13 9.4 T HCa2NaNb2.8Ta1.2O13


9.4 T

0 -200 -400 0 -200 -400


kHz kHz
c) d)
Figure 5.5. 93Nb Static VOCS NMR spectra with fits for HCa2NaNb4O13 at 14.1 (a) and 9.4 T (c) along
with HCa2NaNb2.8Ta1.2O13 at 14.1 (b) and 9.4 T (d). The black line corresponds to the experimental
spectrum and the blue line corresponds to the full calculated spectrum. The other colored lines
correspond to individual fits for each Nb environment. HCa2NaNb4O13 CQ values for individual fits are
as follows: purple = 22.5 MHz; red = 28.7 MHz; green = 29.5 MHz; and orange = 52.1 MHz. For
HCa2NaNb2.8Ta1.2O13 CQ values for individual fits are as follows: purple = 22.3 MHz; red = 25.9 MHz;
green = 27.1 MHz; and orange = 51.1 MHz.

121
Static 93Nb VOCS NMR spectra for RbCa2NaNb4O13 and RbCa2NaNb2.8Ta1.2O13

at 9.4 and 14.1 T shown in Figure 5.6 were fit using 4 sites for both the pure Nb5+ and Ta5+

incorporated compounds. The fit parameters shown in Table 5.2 are consistent with those

seen in the protonated compounds. Three sites showed moderate CQ values in the range of

30-40 MHz and fourth site with a much larger CQ of approximately 50 MHz. These

environments correspond to the exterior and interior Nb octahedra. Furthermore, the

population changes upon incorporation of Ta5+ are consistent with those seen in the

protonated form. As shown in Table 5.2, the population of the interior site drops from

48.2% to 34% indicating a strong preference for Ta5+ at the interior site.

122
RbCa2NaNb4O13 9.4 T RbCa2NaNb2.76Ta1.24O13 9.4 T

0 -200 -400 0 -200 -400


kHz kHz

RbCa2NaNb4O13 14.1 T RbCa2NaNb2.76Ta1.24O13 14.1


T

0 -100 -200 -300 -400 0 -100 -200 -300 -400


kHz kHz

Figure 5.6. 93Nb Static VOCS NMR spectra with fits for RbCa 2NaNb4O13 at 9.4 (top left) and 14.1 T
(bottom left) along with RbCa2NaNb2.76Ta1.24O13 at 9.4 (top right) and 14.1 T (bottom right). The black
line corresponds to the experimental spectrum and the blue line corresponds to the full calculated
spectrum. The other colored lines correspond to individual fits for each Nb environment. For
RbCa2NaNb4O13 CQ values, purple = 33.5 MHz, red = 35.2 MHz, green = 18.8 MHz, and orange = 53.5
MHz. For RbCa2NaNb2.8Ta1.2O13 CQ values, purple = 28 MHz, red = 33 MHz, green = 21 MHz, and
orange = 51 MHz.

123
Table 5.2. NMR fit parameters for RbCa2NaNb4O13 and RbCa2NaNb2.8Ta1.2O13 obtained from VOCS spectra. Error in the last digit is shown in
parentheses.
Compound Site Population CQ η δis0 (ppm) κ Ω (ppm) α (o) β (o) γ (o)
(%) (MHz)
RbCa2NaNb4O13 1 8.2(5) 33.5(1) 0.35(1) -985(1) 0.60(1) 968(2) 9.8(6) 0.0(4) 18.9(6)

2 36.6(9) 35.2(3) 0.38(2) -990(4) 0.62(3) 990(10) 17(3) 3.3(7) 24(3)

3 7.0(4) 18.8(6) 0.40(4) -1040(5) 0.63(2) 640(30) 43(6) 9(3) 11(7)

4 48.2(9) 53.5(4) 0.45(2) -1050(10) 0.06(2) 370(50) 6(1) 11.7(6) 17(4)

RbCa2NaNb2.8Ta1.2O13 1 36(2) 28(1) 0.49(4) -985(20) 0.7(1) 950(30) 3(3) 6(2) 50(20)

2 20(3) 33(1) 0.41(4) -904(10) 1.0(3) 1100(50) 0(10) 14(4) 20(30)


124

3 10(2) 21(1) 0.6(1) -1000(30) -0.7(2) 290(50) 0(15) 0(10) 5(10)

4 34(3) 51(2) 0.41(5) -1060(30) 0.0(2) 390(50) 0(15) 24(3) 3(10)


To ensure physically correct fits were obtained, rather than simply mathematically

correct fits, several methods were employed. Collecting static spectra at multiple fields

ensures fidelity of fits due to varying influence of EFG and CSA parameters at multiple

fields. Furthermore, MQMAS and BRAIN-CP spectra were used to extract various
93
parameters for several Nb environments. The reduced number of interactions of

unknown strength help to ensure the accuracy of static spectra fits. MQMAS spectra for

RbCa2NaNb4O13 and RbCa2NaNb2.8Ta1.2O13 are shown in Figure 5.7. The MQMAS

spectrum for RbCa2NaNb4O13 shows two 93Nb signals. Moment analysis of the two signals

(<v1> = -286 and -259 ppm for sites 1 and 2, respectively) resulted in CQ values of 35.2

and 33.5 MHz, and η values equal to 0.38 and 0.35 for sites 1 and 2, respectively. The

MQMAS spectrum for RbCa2NaNb2.8Ta1.2O13 shows three 93Nb signals. Moment analysis

(<v1> = -182, -251, and -102 ppm for sites 1, 2, and 3, respectively) of the two signals

resulted in CQ values of 28, 33 and 21 MHz, and η values equal to 0.49, 0.41, and 0.6 for

sites 1, 2, and 3, respectively.

125
Figure 5.7. 93Nb 3QMAS 2D spectrum of RbCa2NaNb4O13 (left) and RbCa2NaNb2.8Ta1.2O13 (right)
with double shearing scheme. The horizontal axis corresponds to the isotropic dimension and vertical
axis to the anisotropic dimension.

If Ta5+ had no preference for the interlayer or interior octahedra, the relative 93Nb

populations (Table 5.3) would show no significant change when Ta5+ was present. As seen

in Table 5.3 and Figure 5.5, the relative population of site 4 dropped from 44.6% in the

pure Nb5+ compound down to 32.4% in the Ta5+ containing compound while the other three
93
Nb environments’ relative populations remained unchanged. It should be noted that the

44.6% population is slightly lower than the theoretical 50% population for the interior Nb

environment, but the slight distributions of EFG and CSA with such large magnitudes in

these systems makes exact fitting difficult. If 100% of Ta5+ in the system was in the interior

octahedra, based on the elemental composition the population of site 4 in the 93Nb spectra

should decrease to 28.6%. The population drop to 32.4% indicates Ta5+ has an almost

completely exclusive site preference for the octahedral site associated with the interior

Nb(2) site 4 environment with nearly all Ta5+ in the structure located in this site. The 93Nb

static spectra for the parent compounds RbCa2NaNb4O13 and RbCa2NaNb2.8Ta1.2O13

shown in Figure 5.6 also show substantial population reduction of the larger EFG

126
environment upon Ta5+ incorporation. It should also be noted that the high asymmetry

parameters, η, in the range of approximately 0.4-0.7 support our XRD pattern fits using

orthorhombic unit cells. If the compounds indeed had a tetragonal structure as suggested

originally, the 4-fold symmetry about each Nb site would lead to an η value close to zero

as seen in 93Nb NMR studies of tetragonal RbSr2Nb3O10.169

To confidently assign the structural site associated with the ~50 MHz Nb
93
environment, Nb-1H cross polarization and recoupling NMR experiments were

performed. In the BRAIN-CP experiment, polarization from 1H is transferred to nearby


93
Nb environments. The pulse sequence utilizes WURST pulses that allow for broadband

excitation large enough to cover the Nb environments seen in the static spectra. The

strength of the polarization transfer is inversely proportional to distance, thus only the

interfacial Nb environments would be polarized while the interior Nb sites would remain

unchanged. As seen in Figure 5.8 and Table 5.4, the BRAIN-CP spectrum for the pure

Nb5+ sample shows only signals with moderate CQ values suggesting the 52.1 MHz

environment is indeed associated with the interior octahedral site.

Complementing the cross-polarization data, MQMAS-t2-REDOR spectra also

indicate the moderate CQ environments are associated with the interlayer Nb octahedra. In
93
recoupling experiments, nonzero dipolar dephasing causes attenuation of Nb signals in

close proximity to 1H nuclei. The original MQMAS spectrum (Figure 8a) shows the three

moderate CQ environments as expected, and the signal intensity for all three is distinctly

attenuated when the dephasing pulse on 1H is introduced (Figure 5.9b). The degree of
93
signal attenuation appears to vary somewhat between the three Nb

127
environments suggesting varying Nb-H distances among the exterior octahedra which

implies a degree of disorder is present throughout the structure.

93
Based on the static Nb wideline and 1H-93Nb cross polarization and recoupling

experiments, Ta5+ must have strong preference for the interior octahedral site in the four-

layer compound. Due to Nb5+ and Ta5+ having the same charge and similar size, it was

assumed Ta5+ would either prefer the interlayer octahedral site due to kinetic limitations of

the rapid synthesis or would be randomly distributed throughout the compound. A likely

explanation for the site preference is the difference in electronegativity of Nb 5+ and Ta5+.

As previous studies of Nb5+ and Ta5+ systems have shown, the more ionic Ta-O bonds

inhibit the degree to which octahedra undergo second order Jahn-Teller effects resulting in

TaO6 octahedra exhibiting higher symmetry than their NbO6 counterparts in isostructural

compounds.166, 170-172 In mixed Nb5+/Ta5+ systems, including this one, this manifests itself

as Ta5+ preferentially occupying more symmetric octahedral sites if such sites are

available.165

As mentioned previously, the increased electronegativity of Nb5+ leads to more

covalent Nb-O bonds relative to the Ta-O bonds. In a study of three-layer perovskites

RbLa2-xCaxTi2-xNb1+xO10 (0 < x <2) compounds showing site ordering of -(Ti0.5Nb0.5)O6-

TiO6-(Ti0.5Nb0.5)O6- for x=1 samples, Byeon et al.62 found, using Raman spectroscopy,

that reducing the covalent character of the interior octahedral sites in turn increased the

covalent character of the interlayer octahedral sites. If this behavior is consistent with

mixed Nb5+/Ta5+ systems, the interior octahedral preference of Ta5+ would lead to

increased covalent character of the interfacial Nb(1) site. This increased covalent character

128
would then lead to increased solid acid strength as the increased covalent Nb-O bond would

reduce the charge on the O thus weakening the O-H bond in layered perovskites where

Ta5+ is located exclusively in the interior octahedra compared to pure Nb5+ compounds.

HCa2NaNb4O13 1H-93Nb
BRAIN-CP 9.4 T

0 -50 -100 -150 -200 -250


kHz

Figure 5.8. 1H-93Nb BRAIN-CP spectrum with calculated fit. The black spectrum is the experimental,
blue the full calculated fit, red is the fit of site 3 (28.7 MHz), and green is the fit of site 2 (29.5 MHz).

129
Table 5.3. 93Nb NMR fit parameters for HCa2NaNb4O13 and HCa2NaNb2.8Ta1.2O13 obtained from MQMAS and VOCS spectra. Error in the last
digit is shown in parentheses
Sample Site Population CQ (MHz) η δiso (ppm) κ Ω (ppm) α (o) β (o) γ (o)
(%)
HCa2NaNb4O13 1 6.0(2) 22.5(1) 0.56(2) -1081(5) 0.9(1) 290(10) 73(4) 73(2) 159(4)

2 13.9(4) 28.7(2) 0.65(1) -1113(3) 0.1(1) 875(5) 68(1) 13.4(4) 89(8)

3 35.5(4) 29.5(2) 0.23(2) -1099(2) 0.42(3) 580(10) 12(5) 2.4(9) 71(5)

4 44.6(3) 52.1(3) 0.47(1) -1070(3) 0.37(3) 620(10) 16(2) 9.3(6) 18(3)

HCa2NaNb2.8Ta1.2O13 1 13.7(2) 22.3(2) 0.55(6) -1087(4) 0.91(1) 180(7) 80(15) 74(3) 103(3)

2 31.4(5) 25.9(1) 0.75(1) -1108(3) 0.03(1) 850(4) 73(1) 19.6(7) 77(1)

3 22.5(3) 27.1(3) 0.62(2) -1106(3) 0.63(2) 620(7) 4(2) 3.3(5) 90(2)

4 32.4(2) 51.1(3) 0.46(1) -1057(5) 0.25(2) 650(10) 4(4) 14.8(6) 54(4)


130

Table 5.4 93Nb NMR fit parameters for HCa2NaNb4O13 and HCa2NaNb2.8Ta1.2O13 BRAIN-CP spectra.
Sample Population (%) CQ (MHz) η δiso (ppm) κ Ω (ppm) α (o) β (o) γ (o)

HCa2NaNb4O13 33.3 28.7 0.65 -1100 0.25 850 95.0 18.0 100.0

66.7 29.5 0.23 -1080 0.55 720 10.0 6.0 10.0

HCa2NaNb3.3Ta0.7O13 20.0 25.9 0.62 -1100 0.20 850 95.0 15.0 50.0

80.0 27.1 0.40 -1080 0.0.50 675 10.0 6.0 10.0


Figure 5.9. MQMAS-t2-REDOR spectra for HCa2NaNb4O13 without (left) and with (right) REDOR
sequence active. Numbers correspond to MQMAS sidebands for appropriate sites. Site numbers are
the same as those listed in Table 1.

5.3.3 23Na MAS NMR


The 23Na MQMAS spectrum for the pure Nb compound (Figure 5.10a) indicates

two 23Na environments, consistent with an interior Na site situated between interior Nb(2)

octahedra as well as an exterior Na site situated between interior Nb(2) octahedra and

exterior Nb(1) octahedra. As mentioned previously, if Na is randomly distributed as

suggested in literature reports, there should be a 2 to 1 population for the Na sites. Using

MQMAS fitting and moment analysis as well as fitting the MAS spectra (Figure 5.10b and

Table 5.5), however, revealed two sites with a population ratio of approximately 3 to 1

suggesting Na has a site preference. A rudimentary point charge model developed by

Koller et al.,173 estimating the charge on the various oxygen cations using the calculated

bond valences of the M-O bonds, was implemented to determine the appropriate ratio of

the EFG of the exterior 23Na to the interior 23Na environment. The bond valences, sij, were

calculated using Equation 5.1 where ro is the empirically derived ideal bond length, rij is

131
the bond length in the crystal structure, and B=0.37 is a constant, as demonstrated by

Shannon and Brown174 with bond lengths obtained from Sato et al.35 The charges, qi, on

the oxygen atoms were calculated using Equation 5.2, where a and M are empirically

determined parameters dependent on the number of core electrons of the cation,174 and used

in Equation 5.3 to calculate the resulting EFG tensor, where qe is the charge on the O in

the Na-O bond, x, y, and z are the bond distance components along the various axes, and r

is the bond distance. Finally, the EFG tensor was diagonalized in order to relate Vzz to CQ.

The calculated values of qi for each oxygen atom and resulting V zz for both exterior and

interior 23Na sites are presented in Table 5.6.

𝑟0 − 𝑟𝑖𝑗
Equation 5.1: 𝑠𝑖𝑗 = 𝑒𝑥𝑝 [ 𝐵
]

𝑀
Equation 5.2: 𝑞𝑖 = (∑𝑗 𝑎𝑠𝑖𝑗 )−2

3𝑥𝑖2 −𝑟𝑖2 3𝑥𝑖 𝑦𝑖


Equation 5.3: 𝑉𝑥𝑥 = 𝑒 ∑𝑖 𝑞𝑖 ( ) , 𝑉𝑥𝑦 = 𝑒 ∑𝑖 𝑞𝑖 ( ) , 𝑒𝑡𝑐 …
𝑟𝑖5 𝑟𝑖3

Using the tetragonal crystal structure proposed by Sato et al. as a guide, values for

Vzz were calculated for 23Na environments. Vzz is proportional to CQ and thus the ratio of

Vzz between 23Na environments is equal to the ratio of CQ values between environments.

It was found that the 23Na environment at the interstitial site closer to the interface should

have a CQ value approximately 40% larger than the interior interstitial 23Na environment.

The CQ ratio of the two experimental environments reasonably matches the model

suggesting that the Na+ cation prefers the exterior interstitial site. The MQMAS spectrum

132
and MAS spectrum for the Ta5+ containing compound (Figure 5.10c and 5.10d) show three
23
Na environments with relative populations of 12.8, 19.2, and 68.0 %. Comparing the CQ

and δiso values of HCa2NaNb2.8Ta1.2O13 with the pure Nb compound it appears that Na is

predominately or entirely at the exterior interstitial site based on the increased C Q of all

three 23Na environments in the Ta containing sample. Considering the more ionic nature

of the Ta-O bond, it should be expected that Ca2+ would be preferred over Na+ at the interior

to better compensate for the more negative oxygen environments compared to the exterior

interstitial site. Furthermore, considering the interior interstitial sites are now adjacent to

a mixture of NbO6 and TaO6 octahedra, the observation of two sites is consistent with the

previous determination of interior site preference for Ta5+.

133
HCa2NaNb4O13 23Na
MAS

-1000 -3000 -5000


Hz
a) b)

HCa2NaNb2.8Ta1.2O13
23Na MAS

0 -1000 -2000 -3000 -4000 -5000


Hz
c) d)
Figure 5.9. 23Na MAS and MQMAS spectra for HCa2NaNb4O13 (a and b) and HCa2NaNb2.8Ta1.2O13 (c
and d). The black line corresponds to the full experimental spectrum, the blue line is the full calculated
spectrum, the red line is the individual fit for site 1, the green line is the individual fit for site 2, and
the orange line is the individual fit for site 3. The fit parameters for each site are listed in Table 2.

Table 5.5. 23Na NMR fit parameters obtained from MQMAS and MAS spectra.
Sample Site Pop (%) CQ (MHz) δiso (ppm) η

HCa2NaNb4O13 1 74.1(1.5) 1.93(2) -14.3(2) 0.72(4)

2 25.9(0.7) 1.27(3) -20.8(2) 0.9(1)

HCa2NaNb2.8Ta1.2O13 1 12.8 (4.2) 2.1(1) -14.4(3) 0.78(6)

2 19.2(0.3) 1.9(1) -20.6(2) 0.9(1)


3 68.0(1.2) 1.7(1) -15.5(2) 0.9(1)

134
Table 5.6 Results of EFG tensor calculations using a point charge model for HCa 2NaNb4O13.
Na Site Vzz* (x1024 V/m2) O site Description Bond Length (Å) # of Bonds qi

Interior 0.8 1 Axial Nb(2)-O(1)-Nb(2) 2.73 4 -1.24e

2 Equatorial to Nb(2) 2.75 8 -1.06e

Exterior 1.11 2 Equatorial to Nb(2) 2.98 4 -1.06e

3 Axial Nb(2)-O(3)-Nb(1) 2.76 4 -1.21e

4 Equatorial to Nb(1) 2.5 4 -1.17e

5.4 Conclusions

Our 93Nb and 23Na NMR studies have shown Ta5+ to have a strong site preference

for the interior octahedra in the four-layer RbCa2NaNb4-xTaxO13 systems while Na+ is

believed to be favored at the exterior interstitial site. The use of quadrupolar solid-state

NMR to determine site ordering is a viable alternative to XRD, especially in materials in

which such conclusions cannot be readily made without access to a synchrotron source.

This determination of site preference based on octahedral distortions and electronegativity

can be quite useful in analyzing the results of other work on mixed Nb5+/Ta5+ systems.31,
164
Furthermore, since composition has been shown to have drastic impacts on the

properties of layered perovskites, this observation of site affinity can provide invaluable

insight into the nature of these behaviors and potentially allow for specialized tuning of

properties based on composition.

135
Chapter 6: Microwave Assisted Grafting of Mixed Cation
HCaxSr2-xNb3O10 Compounds with n-Alcohols*
6.1 Introduction

Effective catalytic behavior in the layered perovskite class of compounds requires

access to the protonated environments, typically done through exfoliation into single

nanosheets.63 Sugahara et al.64-69 have pioneered grafting techniques for attachment of

various organic groups into the interlayer of layered perovskites. Sugahara demonstrated

a grafting technique for protonated D-J perovskites to produce derivatives with n-alcohols

up to chain length n=18. Current grafting methods of layered perovskites involve high

pressure heating at 80-150 oC for days to weeks depending on the alcohol involved. The

long reaction time involved inhibits the ability to study organically grafted samples in a

timely fashion. In this current work, a novel microwave irradiation method (Figure 6.1 and

Table 6.1) was developed to reduce the grafting reaction time by 94-97% while maintaining

coverage levels consistent with traditional heating methods. As will be discussed, grafting

leads to a pillaring effect where the interlayer space has greatly increased while potentially

active protonated sites remain.

*Chapter adapted from “Boykin, J. R. and L. J. Smith (2015). "Rapid Microwave-Assisted


Grafting of Layered Perovskites with n-Alcohols." Inorganic Chemistry 54(9): 4177-
4179.”

136
Proton (not to scale)
Sr
NbO6

Alkyl Chain

Microwave Heating
90% Methanol

Microwave Heating
95% Propanol

Microwave Heating
100% Hexanol

Figure 6.1. Illustration of stepwise microwave grafting.

6.2 Experimental

Parent compounds RbCa2Nb3O10 and RbSr2Nb3O10 were synthesized using molten

salt heating of the carbonate and oxide precursors in RbCl salt at 800 oC for 30-45

minutes.29 The hydrated protonated forms were obtained by treating samples with 6M

137
HNO3(aq) at 50 oC for 3 days. The hydrated HSr2Nb3O10 samples (~0.2g for each sample)

were then heated in a domestic microwave oven using a Parr 23 mL microwave acid

digestion bomb with 13.5 ml of 90% methanol using the cycled heating times shown in

Table 6.1. Microwave heating was cycled to prevent overheating of the acid digestion

bomb. After successful methyl grafting, the samples were filtered and washed with

acetone, heated in 9.5 ml of 95% n-propanol, and collected by filtration and washing with

acetone. Samples were then heated in 10.0 ml of 100% n-hexanol with detailed cycles

times shown in Table 6.1, followed by filtration and washing with acetone. HCa2Nb3O10

samples were grafted using the conventional heating method as a way of comparing the

level of grafting with published reports66 as well as with microwave grafted samples. The

HCa2Nb3O10 samples were heated in 90% methanol at 100 oC for 3 days in an acid digestion

bomb then filtered and washed with acetone. This was followed by heating in 100% n-

propanol at 150 oC for 7 days, then filtering and washing with acetone. This was finally

followed by heating in 100% hexanol at 150 oC for 7 days, then filtering and washing with

acetone. Instrumental details of data collection are given in Chapter 3. It should be noted

that while no precise measurement of pressure was obtained, the heating times were chosen

such that pressure reaches a maximum of approximately 500 psi.

138
Table 6.1. Heating Cycles and Reaction Times for Microwave and Conventional Heating Methods
Alcohol Heating Cycle (100% # of Cooling Period Total Reaction
power) Cycles Between Cycles Time
Methanol 5x (20s on 40s off) 10 15-20 min ~4 hours
(microwave)
Methanol 100o C 3 days --- --- 3 days
(conventional)
Propanol 5x (40s on 20s off) 10 15-20 min ~4 hours
(microwave)
Propanol 150o C 7 days --- --- 7 days
(conventional)
Hexanol (microwave) 4x (60s on 15s off) 7 20-25 min ~3.5 hours
Hexanol 150o C 7 days --- --- 7 days
(conventional)

6.3 XRD Results and Discussion

XRD powder patterns were collected for a 2θ range of 2.5-14o and 2.5-60o. The

crystal structure for fully hydrated HSr2Nb3O10 has an orthorhombic space group with

lattice parameters a = 7.808(5), b= 7.82(1), and c = 32.49(2) Ǻ. Upon grafting with n-

alcohols, the changes in the interlayer distance result in changes in the c lattice parameter

as shown in Figure 6.2. The 2.5-14o range contains the reflections for the (001) and (002)

planes associated with the stacking direction in the grafted and hydrated forms,

respectively.

C6/Sr2Nb3O10
C6/Ca2Nb3O10
C1/Sr2Nb3O10
C3/Ca2Nb3O10
C3/Sr2Nb3O10
C1/Ca2Nb3O10
HSr2Nb3O10
HCa2Nb3O10

2.5 4 5.5 7 8.5 10 11.5 13 2.5 4 5.5 7 8.5 10 11.5 13


2θ (o) 2θ (o)
Figure 6.2. (a) XRD patterns for HSr2Nb3O10 (black), C1/Sr2Nb3O10 (red), C3/Sr2Nb3O10 (green), and
C6/Sr2Nb3O10 (blue) synthesized using the microwave irradiation method. (b) XRD patterns for
HCa2Nb3O10 (black), C1/Ca2Nb3O10 (red), C3/Ca2Nb3O10 (green), and C6/Ca2Nb3O10 (blue) synthesized
using a conventional heating method.

139
Despite repeated attempts with modification to all synthetic parameters (including

water content, heating time, number of cycles, and solvent volume), the HCa2Nb3O10

samples never achieved successful grafting beyond the methyl stage. As Figure 6.3

illustrates, XRD patterns for attempted grafting with n-propanol consistently contained

strong reflections associated with the hydrated HCa2Nb3O10 phase and less intense

reflections associated with the C3/Ca2Nb3O10 phase. Furthermore, TGA data show only

20.1 % methyl coverage for C1/Ca2Nb3O10, compared to 35-40 % coverage seen in

C1/Sr2Nb3O10 samples. It was therefore concluded that the Ca form of the three-layer

niobate contains protons so tightly bound that the favorable reaction conditions of

microwave heating are unable to overcome the kinetic barrier of microwave-based grafting.

After the ultimately unsuccessful attempts with microwave grafting of HCa2Nb3O10, the

HSr2Nb3O10 compound was used for attempted grafting.

A large motivation for the potential of HSr2Nb3O10 to be more grafting susceptible

came from the observations of the difficulty in obtaining a fully hydrated HCa2Nb3O10

sample. It has been seen that HCa2Nb3O10 requires a narrow temperature range of ~50-60
o
C during the acid exchanging step in order to achieve a fully acid-exchanged hydrated

state. Furthermore, if HCa2Nb3O10 is not acid exchanged within this necessary temperature

range it is difficult to revert to a hydrated state even with treatment with distilled water

under various conditions including submersion for several days, treatment in a high

humidity environment, or microwave heating with distilled water. The HSr2Nb3O10

compounds however revert to a fully hydrated state simply by being exposed to

atmospheric conditions. In fact, several attempts, made through heating of the HSr2Nb3O10

140
compound to obtain the anhydrous form, have shown that reversion to the hydrated state

occurs within 1-3 hours after the sample has been exposed to the atmosphere even in

situations in which humidity has been measured to be below 25%. Therefore, it was

believed the interlayer gallery of HSr2Nb3O10 is more susceptible to modification than the

HCa2Nb3O10 form. We were quite satisfied to find this assumption was correct and

microwave grafting was incredibly successful in the pure Sr state.

HCa2Nb3O10 Microwave Grafting

3 4 5 6 7 8 9 10 11 12 13 14
2θ (Degrees)

Figure 6.3. XRD patterns for HCa2Nb3O10 (Black), C1/Ca2Nb3O10 (red), and C3/Ca2Nb3O10 (green)
synthesized using microwave grafting. Note the presence of an intense reflection associated with the
HCa2Nb3O10 phase.

The successfully methoxylated sample showed a slight shift of the stacking

reflection to a smaller distance consistent with the decreased size of the interlayer when

water molecules from the hydrated phase are replaced by methyl groups in the C1 stage.

The C3 and C6 samples showed increased stacking distances from a starting point of ~16.2

Ǻ to ~19.8 and ~24.5 Ǻ, respectively. This increase in stacking distance is consistent with

the longer alkyl chains now present in the interlayer gallery. The C3 and C6 samples’ XRD

patterns also show low intensity broad reflections at 2θ = ~5.5O, consistent with a small

141
amount of hydrated phase still being present. Any hydrated phase that remains in the

sample is a small fraction of the final composition.

The changing position of stacking reflections was also seen in the HCa2Nb3O10

samples grafted using conventional heating. Long range XRD patterns were obtained for

C6/Sr2Nb3O10 and C6/Ca2Nb3O10 (Figures 6.4 and 6.5) to confirm three-dimensional

ordering of the compounds remained intact upon alkoxylation. Le Bail175 fits of the long

range patterns performed with the Bruker TOPAS software package showed elongation of

the c-axis consistent with successful grafting while maintaining crystallinity in three

dimensions. Calculated lattice parameters with errors along with particle sizes based on

Scherrer analysis are given in Table 6.2. It should be noted that the grafted samples’ XRD

patterns were fit using a tetragonal P4 unit cell that provides a satisfactory simulated pattern

based on the small number of reflections present and confirms the elongations of the c-

axis, but does not necessarily reflect the true space group of the compounds.

142
Figure 6.4. Long range XRD patterns for HCa2Nb3O10 (top) and HSr2Nb3O10 (bottom). Experimental
patterns are shown in blue. Le Bail fits are shown in red. Blue ticks represent hkl reflections. The
difference pattern is shown in gray.

143
Figure 6.5. Long range XRD patterns for C6/Ca2Nb3O10 (top) and C6/Sr2Nb3O10 (bottom).
Experimental patterns are shown in blue. Le Bail fits are shown in red. Blue ticks represent hkl
reflections for the C6 phase. Black ticks represent hkl reflections for the hydrated phase. The
difference pattern is shown in gray. Listed hkl reflections refer to the C6 phase unless specified
otherwise. *Refers to the minority hydrated phase.

Table 6.2. Le Bail method fit lattice parameters, zero correction, χ2, particle size in stacking direction,
and Rwpa from Le Bail fits for hydrated and C6 samples. Error in the last digit is represented by
parentheses.
Sample Space a (Ǻ) b (Ǻ) c (Ǻ) Particle Zero χ2 Rwp
Group Size (Ǻ)b Corr. (%)
Figure 3
HCa2Nb3O10 Pnma 7.675(1) 7.917(1) 32.333(7) 203(9) -0.0074 1.24 10.49
HSr2Nb3O10 Pnma 7.808(5) 7.82(1) 32.49(2) 350(25) -0.0005 1.18 11.11
Figure 4
C6/Ca2Nb3O10 P4 3.8569(5) 3.8569(5) 24.800(9) 400(11) -0.0032 1.71 7.67
HCa2Nb3O10 P4 7.712(9) 7.716(9) 32.19(4) 80(20) -0.0032 1.71 7.67
C6/Sr2Nb3O10 P4 3.9044(8) 3.9044(8) 24.77(1) 151(5) -0.0042 1.46 9.77
HSr2Nb3O10 P4 7.702(5) 7.702(5) 31.39(3) 140(10) -0.0042 1.46 9.77
a. Rwp refers to the background subtracted R-factor
b. Particle size refers to stacking direction based on (001) and (002) reflections for the grafted and hydrated
compounds, respectively.

144
6.4 TGA and 1H NMR Results and Discussion

While the XRD data indicate the majority of the samples have been successfully

grafted, they unfortunately do not provide quantitative information as to the alkoxyl

coverage levels. To obtain quantitative information about the alkoxyl coverage, 1H NMR

and TGA experiments were performed on the grafted samples. MAS NMR spectra of the

grafted samples contained peaks associated with residual water in the sample, unreacted

acid sites from the Nb-O-H groups, and alkyl protons on the alkoxyl chains as shown in
1
Figure 6.6a. Peak integration of H Hahn Echo176 spectra with varying relaxation

times were used to estimate alkoxyl coverage for each sample by comparing the integral of

the unreacted acid sites to the integral of the alkyl peak as shown in Figure 6.6b. Varying

relaxation times were used to extrapolate back to an echo time of zero and negate the effects

of differing relaxation rates. Methoxylated and C3 samples did not show uniform

exponential decay due to a broad overlapping peak at short relaxation times resulting in a

slightly skewed data set. The 1H NMR results indicate 40% coverage per perovskite unit

for C6 samples.

C6/Sr2Nb3O10 1H NMR with Varying C6/Sr2Nb3O10 Peak Area vs. Echo


Echo DelayTime DelayTime
20
1.4 ppm
Integral Units

-(CH2)5-CH3
15 Peak at 1.4
100 us ppm
800 us 10 Peak at
6.5 ppm 11.4 ppm
11.4 ppm H2O 5
H-O-NbO5

0
ppm 5 0 400tau
15 -5 (μs)800 1200
1
Figure 6.6. (a) H MAS Hahn Echo NMR spectra for C6/Sr2Nb3O10 with varying echo delay times and
(b) results of peak integration as a function of the echo delay time.

145
TGA data shown in Figure 6.7 collected from 30 – 800 oC were collected to

substantiate the NMR results. Water loss was seen up to 110 oC, followed by further mass

loss up to 800 oC. Assuming alcohol loss would occur before condensation of the

compound into a denser phase, the mass loss derivatives were used to determine the

appropriate point to consider alcohol loss complete which was found to be approximately

375 oC. The TGA results are in good agreement with the NMR data and suggest 39.1%

coverage per perovskite unit for the microwave C6 samples. TGA data were also obtained

for the C6/Ca2Nb3O10 compound synthesized using conventional heating; it was found that

the alkoxyl coverage was 45.0%.

100
C6/Sr2Nb3O10 TGA 0.15 C6/Ca2Nb3O10 TGA
100 0.25
Derivative of Weight

Derivative of Weight
98 0.1 98
0.15
Weight %
Weight %

96
96 0.05
%

%
94
0.05
94 0 92
92 -0.05 90 -0.05
100 300 Temp500
(oC) 700 100 300 Temp (500
oC) 700

Figure 6.7. TGA data for C6/Sr2Nb3O10 (above) and C6/Ca2Nb3O10 (below) with temperatures ranging
from 100 to 800 °C. Based on theoretical values of 11.84% and 13.84% for 100% alkoxyl coverage of
the strontium and calcium samples, the alkoxyl coverage was found to be 39.1% and 45.0%,
respectively.

Not only has the total reaction time to obtain the C6 phase been decreased from 17

days to 12 hours, there is no significant decrease in the total alkoxyl group coverage of the

compound. XRD, NMR, and TGA data have shown that microwave irradiation is an

enticing alternative to conventional heating for alkoxylation of layered perovskites. While

using longer chain n-alcohols such as n-octadecanol are not viable due to limited

microwave susceptibility, long chain alcohols could possibly be introduced if more polar

groups are present on the chain to increase the microwave susceptibility. Thorough

146
reviews by Kapp177 and Gabriel et al.178 on microwave heating of organic compounds

discuss in great detail the properties of molecules that affect the ability of a solvent to

absorb microwave energy (referred to as loss tangent), such as the nature of the functional

groups and volume of the molecule. These should serve as excellent resources for those

looking to adopt this microwave assisted grafting technique using other organic species.

6.5 Grafting of Mixed Ca/Sr Compounds

Building off the success of microwave grafting using the HSr2Nb3O10 samples, it

was thought that mixed Ca/Sr forms HCaxSr2-xNb3O10 would potentially exhibit successful

microwave grafting thus imbuing compounds with microwave grafting ability without

substantially altering the composition and thus any other potentially desirable properties.

As shown in Chapter 4, substitution of Ca as low as 10% results in a heterogeneous exterior


93
Nb site as evidenced by fits of Nb NMR spectra for mixed Ca/Sr compounds. It was

observed that based on CQ values, mixed Ca/Sr compounds contained a mixture of Sr-like

environments and Ca-like environments. As discussed in Chapter 4, Ca-like environments

correspond to more highly distorted exterior octahedra when compared to Sr-like

environments. This increased distortion leads to a decreased O-H bond strength and thus

decreased reactivity of the acid site, as evidenced by comparing the catalytic activity of

HSr2Nb3O10 and the more distorted HCa2Nb3O10.52 Due to the presence of less distorted

octahedra in mixed Ca/Sr compounds, it was expected that Sr-like environments would be

able to undergo microwave-assisted grafting beyond C1 stage, even if the more distorted

Ca-like environments cannot. It was initially thought that coverage percentages would

147
gradually increase with increasing Sr content due to the presence of more reactive Sr-like

environments.

To both surprise and delight, it was clear that with as little as 10% Sr substitution,

mixed Ca/Sr samples exhibited grafting ability equal to that of the pure Sr compound. As

shown in Figure 6.8 and Table 6.3, the alkyl coverage for the C1, C3, and, C6 samples of

mixed Ca/Sr compounds exhibited equal or greater coverage than that of the pure Sr

compound.

Table 6.3. Alkyl coverage percentages for C1-C6/CaxSr2-xNb3O10 compounds using microwave
grafting.
Ca Content C1 Coverage (%) C3 Coverage (%) C6 Coverage (%)
0 36.5 29.5 39.1
0.4 43.5 33.6 45.1
0.8 39.2 31.8 43
1 39.7 32.9 43.4
1.5 36.4 33.3 42.8
1.8 36 31.1 44.3
2 20.1 NA NA

Alkyl Coverage vs. Ca content


50
45
40
35
30
25
20
15
10
5
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Ca Content (Formula Units)
Hexyl Coverage Methyl Coverage Propyl Coverage

Figure 6.8. Alkyl coverage percentages for C1-C6/CaxSr2-xNb3O10 compounds using microwave
grafting.

148
The XRD patterns for mixed Ca/Sr grafted samples shown in Figure 6.9 are

remarkably consistent with the results of microwave grafting of the pure HSr2Nb3O10

compounds. The alkyl coverages for C1, C3, and C6 samples shown in Figure 6.8 were

calculated using TGA data alone without additional NMR data due to the consistency seen

between the two methods based on the Sr2Nb3O10 research. The TGA data collection

procedure being less time consuming and less sensitive to the presence of water was

another benefit. Figure 6.10 shows the TGA data for C1-C3/Sr2Nb3O10 samples which are

consistent among all mixed Ca/Sr samples with the exception of the Ca2Nb3O10 samples.

For simplicity, Figure 6.10 has been chosen to represent the general TGA data collected

for all samples. Individual TGA graphs for each sample studied are included in Appendix

B. These results indicate that microwave grafting ability can be imbued with minor

modifications to composition without changing other desirable properties in these D-J

layered niobates.

149
H-C6/Ca0.4Sr1.6Nb3O10 H-C6/Ca0.8Sr1.2Nb3O10
350 1200 400 1400
300 350 1200
1000
C3 Stage C1 Stage
250 300 1000
C6 Stage 800 C3 Stage
250
200 Protonated Stage C6 Stage 800
600 200
150 C1 Stage 600
Protonated Stage
150
400
100 100 400
50 200 200
50
0 0 0 0
2 3 4 5 6 7 8 9 10 11 12 13 14 2 3 4 5 6 7 8 9 10 11 12 13 14
2θ (Degrees) 2θ (Degrees)

H-C6/Ca1.0Sr1.0Nb3O10 H-C6/Ca1.2Sr0.8Nb3O10
600 1400 500 1400
450
500 1200 1200
Protonated Stage 400 Protonated Stage
1000 350 1000
400 C1 Stage C1 Stage
800 300 800
C3 Stage C3 Stage
300 250
C6 Stage 600 200 C6 Stage 600
200 150
400 400
100 100
200 200
50
0 0 0 0
2 3 4 5 6 7 8 9 10 11 12 13 14 2 3 4 5 6 7 8 9 10 11 12 13 14
2θ (Degrees) 2θ (Degrees)

H-C6/Ca1.5Sr0.5Nb3O10 H-C6/Ca1.8Sr0.2Nb3O10
3000 900 500 1000
800 450 900
2500 400 800
C1 Stage 700 C6 Stage
2000 600
Protonated Stage
350 C1 Stage
700
300 600
C3 Stage 500 Protonated Stage
1500 250 500
C6 Stage 400 C3 Stage
200 400
1000 300 150 300
200 100 200
500
100 50 100
0 0 0 0
2 3 4 5 6 7 8 9 10 11 12 13 14 2 3 4 5 6 7 8 9 10 11 12 13 14
2θ (Degrees) 2θ (Degrees)

Figure 6.9. XRD patterns for H-C6/CaxSr2-xNb3O10 compounds grafted using the microwave heating
method.

150
As mentioned previously, alkyl group mass loss in TGA data was determined using

the derivative of weight % to isolate alkyl loss from condensation of the layered structure.

As illustrated in Figure 6.8, the C3 stage for all successfully grafted samples shows the

peculiar trend of decreasing in alkyl coverage by 5-10% from the C1 stage while the

subsequent C6 stage increases by 5-10%. Further analysis of the TGA data for C3

compounds provides an explanation for this phenomenon. While TGA graphs for C1 and

C6 compounds show one distinct mass loss event in the 275-400 oC range, the C3 graphs

show two mass loss events. The mass loss event in the C6 stage occurs at ~100 oC higher

temperature than the C1 stage due to the longer alkyl chain present. Looking at the C3

stage, the first mass loss event occurs at a temperature consistent with that of the C1 stage

while the second mass loss event occurs at a temperature between the mass loss

temperatures of the C1 and C6 stages and can be assigned to propyl group loss. One

possible conclusion is that the C3 samples contain both methyl and propyl sites, which

would also explain the presence of two mass loss events in C3 samples. Unfortunately, the

current TGA data are insufficient to accurately test this hypothesis.

Figure 6.10. TGA graphs for C1 and C3/Sr2Nb3O10 samples. All mixed Ca/Sr samples show the same
general shape of graph with the only changes being the magnitude of mass loss.

151
It is believed that the presence of even small amounts of Sr allows for alteration of

exterior Nb sites beyond those immediately bonded to Sr cations creating an environment

suitable for microwave grafting. This is supported strongly by the results of grafting

coverage in mixed Ca/Sr samples. Despite varying Ca contents, coverage in all stages is

rather consistent with the expected progressions stated previously. Furthermore, this is

consistent with the NMR results of Chapter 4 which showed that the ratios of Sr-like to

Ca-like 93Nb environments was not equal to the atomic ratio of Sr to Ca. Therefore, the

presence of Sr is clearly having an effect extending to nearby sites containing only Ca,
93
consistent with observed changes in CQ of Ca-like environments seen in the Nb NMR

data in Chapter 4.

6.6 UV-visible Results and Discussion

Having successfully applied the microwave assisted grafting technique to mixed

Ca/Sr compounds, UV-visible spectroscopy was used to study the effect of grafting on

band gap. An ISR-3100 integrating sphere attachment was used in conjunction with a

Perkin Elmer Lambda 35 UV-visible spectrophotometer to obtain diffuse reflectance

spectra. Barium sulfate was used as a standard. Tauc plots were obtained using the method

first developed by Tauc, Davis, and Mott179-181 based on the following expression:

Equation 6.1: (ℎ𝜈𝛼)1/𝑛 = 𝐴(ℎ𝜈 – 𝐸𝑔 )

Where h = Planck’s constant, ν = frequency, α = absorption coefficient, A = proportional

constant, Eg = band gap, and n denotes the nature of the transition (equal to ½ in this work

152
for the direct allowed transition). The acquired spectra were converted to the Kubelka-

Munk function with the vertical axis converted to the quantity F(R∞) (F(R∞) = (1- R∞)2/2

R∞) which is proportional to the absorption coefficient. The expression in equation 2.1

thus becomes:

Equation 6.2: (hνF(R∞))2 = A(hν – Eg)

Using the Kubelka-Munk function, the value of (hν – (hνF(R∞))2) is plotted with hν (with

values in eV) on the horizontal axis and (hνF(R∞)2 on the vertical axis. To calculate the

band gap, Eg, a line is drawn tangent to the point of inflection on the curve and the hν value

at the intersection of the tangent line and the horizontal axis is thus equal to Eg.

As shown in Figure 6.11, the direct band gap undergoes a substantial shift upon

grafting with Eg = 3.28 eV for HSr2Nb3O10, 3.64 eV for C1/Sr2Nb3O10, 3.55 eV for

C3/Sr2Nb3O10, and 3.46 eV for C6/Sr2Nb3O10. The value of 3.28 eV for HSr2Nb3O10 is in

line with literature reported values182 indicating the band gap determination procedure was

applied correctly. It has been shown that compositional differences of D-J niobates lead to

changes in band gap such as HSr2Nb3O10 being reported with Eg = 3.30 eV and

HCa2Nb3O10 being reported with 3.50 eV.182 The results shown here, however, are the first

reported instances of band gap modification through organic grafting. While investigation

into band gap engineering through organic grafting methods lay outside the scope of the

work presented here, the effect of grafting on band gap of layered perovskites is most

certainly deserving of further study.

153
HSr2Nb3O10 Tauc Plot
1.7

(hνF(R∞))2
1.1
Experimental
0.5
Fit

-0.1
2 2.5 3 3.5 4 4.5
a) Energy (eV)

C1/Sr2Nb3O10 Tauc Plot


5
(hνF(R∞))2

3
Experimental
1 Fit

-1 2 2.5 3 3.5 4 4.5


b) Energy (eV)

C3/Sr2Nb3O10 Tauc Plot


7
(hνF(R∞))2

3 Experimental

1 Fit

-1 2 2.5 3 3.5 4 4.5


c) Energy (eV)

C6/Sr2Nb3O10 Tauc Plot


2.5
(hνF(R∞))2

1.5
Experimental
0.5 Fit

-0.5 2 2.5 3 3.5 4 4.5


d) Energy (eV)
Figure 6.11. Tauc plots for H-C6/Sr2Nb3O10 (a-d, respectively) microwave grafted samples showing
substantial shift in band gap upon grafting.

154
6.7 Conclusions

The substantial reduction in total reaction time using microwave heating compared

to conventional heating clearly increases the viability of further research on grafting of

layered perovskites. In fact, two other research groups have already begun using

microwave assisted grafting for their work. Wiley et al.183 have used microwave assisted

grafting to successfully graft the two layer HLaNb2O7 compound while also confirming

our results of HCa2Nb3O10 microwave grafting in that the Ca form is unable to graft past

the C1 stage. Wang et al.184-185 have applied the microwave grafting method to the post-

synthetic modification of the Aurivillius phase Bi2SrTa2O9 using amines and diamines

further illustrating the broad applicability of this method. Building on the success of this

grafting work, the next stage of research was to investigate the ability of these grafted

compounds to be exfoliated into nanosheets which will be discussed in the following

chapter.

155
Chapter 7: Exfoliation of Grafted Niobates Using Organic
Solvents

7.1 Introduction

As previous chapters have discussed, the ion-exchange ability of D-J layered

perovskites is one of the properties that has stimulated research interest in recent years.

This ion-exchangeability is key to currently employed exfoliation techniques. The

conventional method of D-J layered niobate exfoliation relies on the intercalation of bulky

bases such as TBAOH after interlayer proton exchange of parent compounds.36, 38, 52, 186-189

The intercalated base increased the interlayer distance thus overcoming forces holding the

layers together and resulting in exfoliation of layers into single nanosheets. This method,

while successful, requires the use of an aqueous solvent and is rather time consuming with

full intercalation only being achieved after treatment for seven days. Furthermore, removal

of TBA cations requires a dropwise addition of dilute acid over the course of several hours,

and if the rate is too fast or too much of an excess of acid is added, sheets will restack and

revert to the initial protonated state. This difficult method of exfoliation is a hindrance in

further research into the properties and applications of niobate nanosheets.

Organic solvents have been used successfully for exfoliation of many layered

materials with a graphite like structure such as metal chalcogenides190-195 (e.g. MoS2, WS2,

GaS, and GaSe to name a few). Typically, with an appropriate solvent, sonication alone

was sufficient to produce colloidal suspensions of exfoliated materials. The efficiency of

the exfoliation solvent was found to be partially dependent on its Hansen solubility

parameters (HSP), a semi-empirical explanation of dissolution behavior. Three HSP values

156
are used: δD, δP, and δH, corresponding to the strength of dispersive, polar, and hydrogen-

bonding forces of the solvent with itself. In work from Zhou et al.159 it was found that

having a close HSP match between the solvent and the nanomaterial for all three HSP

values was key to successful exfoliation. This method of exfoliation in organic solvents

has been found to lack the issues of structural deformations, sensitivity to environmental

conditions, and layer reaggregation that are persistent problems with ion intercalation

methods.39, 182, 196 Up to this point, however, organic solvent exfoliation has not been

applied to layered perovskites due to the strong electrostatic forces holding perovskite

layers in place. Herein, a new method of D-J layered niobate exfoliation using organic

solvents and grafted niobates is presented.

After the success of microwave grafting using n-alcohols, it was thought that the

grafted compounds could be exfoliated using a novel method of treatment with organic

compounds. As the results in Chapter 6 show, stacking distance increases 16.2 Å to ~25

Å going from HSr2Nb3O10 to C6/Sr2Nb3O10 due to substantial expansion of the interlayer

gallery. It was therefore believed that the electrostatic forces holding the sheets together

were sufficiently weakened that treatment with an appropriate solvent would allow for easy

exfoliation. Due to the presence of alkyl chains in the interlayer gallery, introducing an

organic solvent into the interlayer would lead to weakening of the dispersion forces holding

sheets together without the need for modification to the surface sites as is the case with

TBAOH exfoliation. A variety of different solvents and synthetic techniques were

employed to test this hypothesis. As the following sections will show, grafted layered

157
niobates are well-suited for suited for organic solvent exfoliation using relatively rapid

synthetic methods.

7.2 Experimental

7.2.1 Sample Overview

Table 7.1 includes summaries of the exfoliation procedures employed for the 15

samples to be discussed in this chapter. It should be noted that each sample number is

representative of a unique procedure that was repeated multiple times to confirm results.

The synthetic variables studied were composition of the parent compound, solvent choice,

heating, solvent choice, sonication (before and after heating), sedimentation method, and

methods of extracting nanosheets out of solution. Instrumental details for synthetic

procedures and XRD, SEM, EDS, UV-visible, and TGA characterization methods can be

found in Chapter 3. Preparation methods for data collection were highly sample dependent

and are thus discussed as the data are presented.

158
Table 7.1 Summary of exfoliation procedure by sample. More detailed descriptions of the exfoliation procedures are also in the text.
Initial Final SNb
# Composition Solvent Heatinga Centrifuge Notes
Sonication Sonication Treatment

1 C6/Sr2Nb3O10 Cyclohexane 30 min X X X X No colloidal suspension

6000 rpm Air Dry


2 C6/Sr2Nb3O10 Pyridine 30 min X X No suspension
45 min Dropwise
6000 rpm
3 C6/Sr2Nb3O10 5M2Hc 30 min X X White suspension
45 min
6000 rpm Air Dry
4 C6/Ca2Nb3O10 5M2H 60 mind X X White suspension
45 min Dropwise
Dry
6000 rpm
5 C6/Sr2Nb3O10 5M2H X 10x X Dropwise Strongly yellow suspension
45 min
at 80 oC
6000 rpm Air Dry
6 C6/Sr2Nb3O10 5M2H X 8x 60 min Yellow suspension
159

45 min Dropwise
150C 3d
More pale yellow compared to
7 C6/Sr2Nb3O10 5M2H X Steel X X X
microwaved suspensions
Bomb

8 C3/Sr2Nb3O10 2Pe 60 min 10x 60 min X X Slight yellow suspension

Consistent with pure Sr form in Sample


9 C3/Ca1Sr1Nb3O10 2P 60 min 10x 60 min X X
8
6000 rpm Air Dry Consistent with pure Sr form in Sample
10 C6/Ca0.25Sr1.75Nb3O10 5M2H 180 min 10x 180 min
45 min Dropwise 6
Settled faster than 10 after initial
6000 rpm Air Dry
11 C6/Ca0.25Sr1.75Nb3O10 2P 180 min 10x 180 min sonication (no Tyndall effect), slight
45 min Dropwise
yellow color after heating
Table 7.1 Cont’d.
Initial Final SN
# Composition Solvent Heatinga Centrifuge Notes
Sonication Sonication Treatment
6000 rpm 45 Freeze 3.0 mL of SN mixed with 6.0 mL t-BuOH
12 C6/Sr2Nb3O10 5M2H 60 min 10x 60 min
min Dryingf before freeze drying
Dried
6000 rpm 45 SN was distilled at 140-144 oC. Distillate
13 C6/Sr2Nb3O10 5M2H 60 min 10x 60 min Dropwise
min was clear and colorless
under N2
5.0 mL SN mixed with 5.0 mL t-BuOH
6000 rpm 45 Freeze then dried under air stream with regular
14a C3/Sr2Nb3O10 2P 60 min 10x 60 min
min Drying addition of extra t-BuOH until sample had
no scent of 2P
5.0 mL SN from 14a mixed with 2.0 mL
Freeze ethanol followed by 5.0 mL H2O (thin film
14b C3/Sr2Nb3O10 H2O/Ethanol X X X X
Drying at interface). Aqueous phased was used for
freeze drying.
6000 rpm 45 Sediment from 14a was used as starting
14c C3/Sr2Nb3O10 2P 60 min 10x 60 min Air Drying
160

min material.
5.0 mL SN mixed with 5.0 mL t-BuOH
6000 rpm 45 Freeze then dried under air stream with regular
15a C6/Sr2Nb3O10 5M2Hg 60 min X X
min Drying addition of extra t-BuOH until sample had
no scent of 5M2H
5.0 mL SN from 15a mixed with 2.0 mL
Freeze
15b C6/Sr2Nb3O10 H2O/Ethanol X X X X ethanol followed by 5.0 mL H2O.
Drying
Aqueous phase was used for freeze drying.
6000 rpm 45 Rapid Sediment from 15a was used as starting
15c C6/Sr2Nb3O10 5M2H 60 min 10x 60 min
min Drying material.
a) Heating cycle was 5x(45s on 15s off) for both 2P and 5M2H samples.
b) SN = Supernatant
c) 5M2H = 5-methyl-2-hexanone
d) Initially stirred for 3 days in N2 atmosphere.
e) 2P = 2-pentanone
f) Suspension was brought to -25 oC and freeze dried at ~1 mmHg pressure with no additional heat added to the system
g) Organic phase in top layer was evaporated using a stream of air. Thin film remained on top of aqueous phase
h) Sample was prepared using C6/Sr2Nb3O10 in 5M2H suspension made 6 months earlier.
7.2.2 Solvent Choice

The premise of this work relies on the ability to weaken the intermolecular forces

between grafted alkyl chains, and thus the need to study the effects of differing solvents is

self-evident. Solvents in samples 1 and 2 were chosen due to their laboratory ubiquity. 5-

methyl-2-hexanone (5M2H) and 2-propanone were chosen based on their Hansen

Solubility Parameters197 (Table 7.2). Since the grafted alcohols were no longer in their

original form due to the removal of the C-O-H bond and the formation of the C-O-Nb bond,

using the parameter values for the initial alcohol was not ideal. However, since the

propoxyl substructure remains present, certain assumptions can be made. Modification of

the O-H bond would not significantly alter the dispersion forces or dipole-dipole

interactions, but rather it would reduce hydrogen bonding. Therefore, it was believed that

a solvent with similar δD and δP values to that of the initial alcohol, but with a low δH value

would be appropriate for exfoliation. Thus, 5M2H and 2P were the solvents of choice for

C6 and C3 compounds, respectively. It should be noted that there are readily available

solvents other than 2P with closer solubility matches to the propoxyl group such as 1,1-

dichloroethane and chloromethane (with δD, δp, and δH values equal to 16.6/15.3, 8.2/6.1,

and 0.4/3.9 MPa1/2, respectively197), but the similarity of structure between 2P and 5M2H

allows for easier comparisons between samples.

Table 7.2 Hansen Solubility Parameters for exfoliation solvents and alcohols used during grafting of
parent compounds. Values are given in units of MPa 1/2. Values taken from Hansen.197
Name δD δP δH
1-hexanol 15.9 5.8 12.5
1-propanol 16 6.8 17.4
5M2H 16 5.7 4.1
2-pentanone 15.5 10.4 6.7

161
7.2.4 Sonication

Variable sonication steps were present both before and after heating steps. Initial

sonication was thought to reduce particle size and prepare particles for easier interlayer

solvent introduction. As subsequent sections will illustrate, the initial sonication itself

produced exfoliated materials (albeit not in substantial amounts) and is a vital step in the

optimized exfoliation procedure. The final sonication step was thought to assist in the

exfoliation of stacked sheets that may have solvent in the interlayer, but had yet to break

apart. The efficacy of this final sonication step is ambiguous, but is included due to the

effectiveness of the initial sonication and its ease of implementation.

7.2.5 Sedimentation

Sedimentation behavior is highly dependent on particle size in both lateral and

stacking dimensions.198-200 The most basic sedimentation method was to let samples sit

untouched on the benchtop letting the largest particles settle due to gravitational forces

alone. Centrifugation of samples at 6000 rpm (equivalent to 4303 g-force) for 45 minutes

is the traditional technique of separating exfoliated and unexfoliated nanosheets using the

TBAOH method198, 201 and was thus employed in this work for the same reason. The

supernatant (SN) was carefully pipetted away for further use. The g-force during

centrifugation should be sufficient such that only exfoliated single nanosheets remain in

solution

162
7.2.6 Extracting Nanosheets Out of Solution

Lyophilizing (or freeze-drying) is the preferred method of collecting single

nanosheets out of solution using the TBAOH method. This technique involves freezing a

suspension and placing it under vacuum such that the system pressure allows for

sublimation until all solvent has been removed and only nanosheets remain. Simply

evaporating the liquid solvent inevitably leads to sheet restacking. The freezing points of

5M2H and 2P are too low (-73.9 and -76.8 oC, respectively) to be used alone in the

lyophilizing system in our lab. Tert-butanol (t-BuOH, m.p. 25.8 oC) has been applied

previously as a co-solvent202-203 to increase solution melting point and was used with

limited success in this work. Attempts at transferring sheets into an aqueous phase were

also tested and will be discussed in greater detail in subsequent sections.

Many suspensions and SN’s were treated by evaporation of the solvent to obtain

materials for characterization. As the following sections will show, solvent evaporation

inevitably leads to restacking of exfoliated sheets. This restacking, however, is disordered

in nature and can be observed in XRD patterns and SEM images, thus serving as evidence

of the presence of exfoliated sheets in solution. The method of evaporation was varied and

will be discussed in greater detail in later sections.

7.3 Visual Observations and Generation of Colored Suspensions

Figures 7.1-8 show sample images at different synthetic stages. These simple

visual observations provide great information regarding the nature of these suspensions.

Sonication alone in an appropriate solvents was the simplest method of producing distinct

163
colloidal suspensions. The image in Figure 7.1 of C6/Ca2Nb3O10 (Sample 4) after stirring

in 5M2H for 3 days in a N2 atmosphere followed by 60 minutes of sonication shows that

after one week untouched, particles are still suspended in solution. This sample showed

little change after sitting untouched for over 2 months. The SN following centrifugation

was clear and colorless with no observed Tyndall Effect. C6/Sr2Nb3O10 (Sample 3)

sonicated 30 minutes in 5M2H without prior stirring also produced a colloidal suspension,

but to a substantially lesser degree than Sample 4. Solvent choice was found to be vital for

producing colloidal suspensions after sonication as seen in Figure 7.2 showing Sample 10

(C6 phase in 5M2H) and Sample 11 (C6 phase in 2P) after 180 minutes of sonication.

Figure 7.1. Image of C6/Ca2Nb3O10 in 5M2H (Sample 4) after sitting unmoved for one week.

164
Figure 7.2. Image of C6/Ca0.25Sr1.75Nb3O10 in 2P (Left – Sample 11), and in 5M2H (Right – Sample 10)
30 minutes after initial sonication. Note that particles in Sample 11 are settling at a much faster rate
than those in Sample 10.

The addition of a microwave heating stage after initial sonication produced colored

suspensions. Figures 7.3 and 7.4 show C6 and C3/Sr2Nb3O10 (Samples 5 and 8,

respectively) in 5M2H and 2P, respectively, show the distinct color generated after

microwave heating. C6 samples in 5M2H consistently developed strong yellow colors

after microwave heating while C3 samples in 2P samples developed slight yellow tints.

Potential causes of this dramatic color change will be discussed later in this section. The

suspensions remain colored after sitting for several weeks.

165
Figure 7.3. Image of C6/Sr2Nb3O10 in 5M2H (Sample 5) shortly after microwave heating (left) and after
sitting untouched for 1 week (right). The observed Tyndall Effect is indicative of a colloidal suspension
that does not settle out with gravitational force alone.

Figure 7.4. Images for C3/Sr2Nb3O10 in 2P (Sample 8) after sitting untouched for 2 weeks (Left) and 2
months (Right). The colloidal suspension is clearly stable after 2 months even despite approximately
50% of the solvent having evaporated. A slight yellow tint can be observed in the sample after 2 weeks,
but is ambiguous after 2 months.

Sample 15a (C6/Sr2Nb3O10 in 5M2H) was initially prepared with 60 minutes initial

sonication, microwave heating 10x, and 60 minutes sonication. The data presented herein

were collected on the suspension 6 months after initial preparation at which time the sample

was sonicated 60 minutes. 5M2H was periodically added to the suspension to prevent

complete solvent evaporation. As seen in Figure 7.5, the initial yellow color had faded

substantially in subsequent months with only a faint tint remaining. The suspension,

166
however, still clearly contained colloidal particles consistent with sheet exfoliation. The

SN after centrifugation (Figure 7.6 – left) was used to collect suspended nanosheets

through extraction into an aqueous phase. An aliquot of the SN was first thoroughly mixed

with ethanol (miscible in 5M2H) followed by mixing with water (immiscible in 5M2H)

leading to a substantial cloudiness being produced in the aqueous phase (Figure 7.6-right),

possibly due to the presence of particles extracted into the aqueous phase. The

effectiveness of this technique is discussed in section 7.7.

Figure 7.5. Image of C6/Sr2Nb3O10 in 5M2H (Sample 15a) having sit for 5 hr after initial sonication.
Note the yellow color seen in Sample 12 from which it was prepared is barely present, if present at all.

167
Figure 7.6. Image of C6/Sr2Nb3O10 in 5M2H after centrifugation (Sample 15a – Left) and after mixing
with ethanol and H2O (Sample 15b – Right). The organic layer in Sample 15b is on top and aqueous
layer on bottom. Note the substantial cloudiness present in the aqueous layer in Sample 15b.

As section 7.4 will show, since the sediment following exfoliation treatment still

contained grafted material, it was believed that the sediment could undergo a second

exfoliation treatment to produce more nanosheets from the parent sample. Samples 14c

and 15c (C3/Sr2Nb3O10 in 2P and C6/Sr2Nb3O10 in 5M2H) were prepared using the

sediment of samples 14a and 15a, respectively. Figure 7.7 shows the SNs of 14c and 15c

after centrifugation at which point their appearance is consistent with the initially prepared

samples. It should be noted that 15c had a strong yellow color that had faded in the original

sample. Furthermore, Figure 7.8 shows that the Tyndall Effect was observed in the SN

after centrifugation. In all previous samples the Tyndall Effect was observed in the initial

suspensions, but was not observed after centrifugation due either to decreased sizes of

particles present or lower concentration of particles. This would suggest the first treatment

of the sample primed the sediment for later exfoliation

168
Figure 7.7. Image of C3/Sr2Nb3O10 in 2P 2nd exfoliation SN (Sample 14c) and C6/Sr2Nb3O10 in 5M2H
2nd exfoliation (Sample 15c). Note the yellow color in Sample 15c returned after microwave heating

Figure 7.8. Image of C6/Sr2Nb3O10 in 5M2H SN (Sample 15c) with laser beam passing through.
Observance of the Tyndall effect is evidence of a colloidal suspension. This is the first instance in which
the Tyndall Effect was observed after centrifugation.

Considering the role of the solvent in this new exfoliation method was initially

considered to be only as a medium in which to suspend nanosheets, it is tempting to simply

brush aside the emergence of color in suspensions. The origin of this color, however, is

169
clearly related to the properties of layered niobates, and the absence of any other reported

observation of such change in both published reports and informal discussions with other

researchers in the field indicate that new insights into the nature of these materials are ready

to be illuminated.

Considering the oft-touted potential of niobate nanosheets as catalytic materials,

the possibility of a chemical reaction generating a colored product needed to be addressed.

UV-visible spectra for 5M2H untreated, a control sample of 5M2H sonicated and

microwave heated, and C6/Sr2Nb3O10 in 5M2H are shown in Figure 7.9. The untreated

and control samples show no significant differences while the nanosheet sample absorbs

into the visible spectrum consistent with the observed yellow color. The diffuse reflectance

UV-visible spectrum of the C6/Sr2Nb3O10 parent compound (Figure 7.10) shows

absorption at lower wavelength than 5M2H confirming a new component in present in the

sample, though its identity is not yet clear.

5.9

4.9

3.9
Absorbance

2.9

1.9

0.9

-0.1
200 250 300 350 400 450 500 550 600
Wavelength (nm)
5M2H Control nano no butanol

Figure 7.9. UV-vis spectra for untreated 5M2H (Blue), 5M2H control sample (Orange), and Nanosheet
suspension (yellow).

170
C6/Sr2Nb3O10
2.5

1.5
K-M

0.5

0
250 300 350 400 450 500 550 600
Wavelength, nm
Figure 7.10. UV-vis spectrum of C6/Sr2Nb3O10 parent compound.

Figure 7.11 shows comparisons of IR spectra for untreated solvents and the

resulting colored SNs. The suspensions of exfoliated niobates produced no new IR signals

nor do solvent signals show any shifting outside of instrumental error. Furthermore,

comparisons of 1H NMR spectra (Figure 7.12) of untreated solvent, control samples

(solvent treated without the presence of niobates), and colloidal suspension SNs show no

signals except for those from the solvent. The lack of significant differences in IR and

NMR spectra of these samples, however, does not conclude the absence of a chemical

reaction due to limits of detection, but rather provides no additional evidence as the to the

nature of the yellow color.

171
IR Spectra
100
95
90

5M2H/t-BuOH 85

%T
2P 80

C6 in 5M2H/t-BuOH (Sample 15a) 75

C3 in 2P (sample 14a) 70
65
60
3800 3400 3000 2600 2200 1800 1400 1000 600
Wavenumber, cm-1
Figure 7.11. IR spectra for 5M2H/t-BuOH mixture untreated (Blue), 2P untreated (orange),
C6/Sr2Nb3O10 in 5M2H/t-BuOH mixture (grey), and C3/Sr2Nb3O10 in 2P (yellow).

Figure 7.12. 1H NMR spectra for 5M2H untreated, 5M2H control sample, and C6/Sr 2Nb3O10 in 5M2H
SN.

172
7.4 XRD Results

As previous chapters have shown, XRD is an incredibly useful tool for structural

characterization of layered perovskites. Stacking reflections seen in patterns collected with

a 2θ range of 1.5-14o are particularly useful in this work due to their shifting positions

depending on the composition of the interlayer gallery (see Chapter 6 for a more detailed

discussion) and thus relate to the stacking motif of the sample (i.e. ordered vs. disordered

stacking, particle size, presence of alkyl groups, etc.…) A summary of short-range XRD

pattern results including stacking distance, corresponding phases, and mean particle size in

the stacking dimension can be found in Table 7.3. Mean particle size relates to particle

size in the ordered domain and was calculated using Scherrer’s equation. Long range

patterns were also employed to both confirm the perovskite structure is intact after

exfoliation as well as to study the long-range order of the samples. XRD patterns were

collected for the sediment after the initial preparation of colloidal suspensions, suspensions

dried onto glass slides after settling (and before centrifugation), and for solids collected

from SNs after centrifugation. Several methods were employed for collecting solids from

SNs and will be discussed in greater detail in subsequent sections.

173
Table 7.3. XRD Stacking Reflections Summary
Sample Reflection Position Stacking Distance (Å) Phase Mean Particle
(2θ Degrees) Size (Å)
13 - SN 5.65 15.63 Protonated 140
14 - SN 5.70 15.50 Protonated 160
8 - SN 5.68 15.55 Protonated 160
5 - SN 5.64 15.66 Protonated 40
9 - Sus day of 4.30 20.54 C3 80
5.75 15.36 Protonated 140
9 - Sus 2 days 6.00 14.72 Protonated 160
7 - FS Sus 3.94 22.41 C6 110
5.74 15.39 Protonated 140
7 - SS Sus 5.74 15.39 Protonated 140
5 - Sus 3.52 25.09 C6 180
4 - Sus 3.54 24.95 C6 80
5.76 15.34 Protonated 200
15c Sed 5.80 15.23 Protonated 140
14c Sed 4.80 18.40 C3 70
5.76 15.34 Protonated 200
6 Sed 3.02 29.24 C6 90
5.58 15.83 Protonated 110
5 Sed 3.70 23.87 C6 110
5.65 15.63 Protonated 140
4 Sed 3.12 28.30 C6 190

7.2.3 Heating

The results of the grafting work discussed in Chapter 6 clearly illustrate the

important role of heating in introducing organic groups to the interlayer gallery. Therefore,

it was thought that additional thermal energy would allow solvent to enter the interlayer

gallery with greater ease. Microwave heating was employed using a heating cycle of

5x(45s on 15s off) repeated a varied number of time with 15-20 minute cooling intervals

between heating cycles. It should be noted that while 2P has a higher microwave

susceptibility than 5M2H,177 the same heating cycle was employed with both solvents.

174
This was done to prevent overheating of the acid digestion vessel and thus 2P will reach

higher temperatures than 5M2H during microwave heating.

7.4.1 Sediments of Suspensions

Figures 7.13-17 show XRD patterns for Samples 4 (C6/Ca2Nb3O10 in 5M2H), 5

(C6/Sr2Nb3O10 in 5M2H), 6 (C6/Sr2Nb3O10 in 5M2H), 14c (C3/Sr2Nb3O10 in 5M2H), and

15c (C6/Sr2Nb3O10 in 5M2H), respectively. The presence of sharp and intense stacking

reflections associated with the C6 phase in Figures 7.13 and 7.14 strongly suggest the

suspension sediments can be reused in another exfoliation attempt because of the remaining

presence of alkyl groups in the interlayer of unexfoliated sheets. Furthermore, Figure 7.15

shows the XRD pattern for a C6/Sr2Nb3O10 in 5M2H suspension that, due to benign

neglect, sat untouched for over one year after it was prepared (the sample was never

centrifuged). During this time, the 5M2H solvent slowly evaporated until only dry powder

remained on the bottom of the container. The presence of a C6 stacking reflection is a

testament to the relative stability of the C6 phase (whereas C3 phases quickly revert to

hydrated states within days) and suggests the C6 phase is better suited for exfoliation. This

is also supported by the XRD patterns in Figures 7.16 and 7.17 for sediments of 14c and

15c, respectively. The presence of a C3 stacking reflection in 14c and the absence of a C6

stacking reflection in 15c, combined with visual observations (see Figures 7.7 and 7.8) is

indicative of a larger fraction of the total sample having been exfoliated.

175
C6/Ca2Nb3O10 Sediment Sample 4

2 4 6 8 10 12 14
2θ (Deg)

Figure 7.13. XRD pattern for C6/Ca2Nb3O10 in 5M2H sediment collected from the bottom of the sample
container (Sample 4).

C6/Sr2Nb3O10 in 5M2H Sediment Sample 5

2 8 14 20 26 32 38 44 50 56
2θ (Deg)

Figure 7.14. XRD patterns for C6/Sr2Nb3O10 in 5M2H (Sample 5) slow air-dried supernatant (top) and
the powder sediment (bottom).

176
C6/Sr2Nb3O10 in 5M2H (Sample 6) Suspension Dried Over
1 year

2 4 6 8 10 12 14
2θ (Deg)

Figure 7.15. XRD pattern of C6/Sr2Nb3O10 in 5M2H (Sample 6) after it had sat on the benchtop
untouched for over 1 year. All solvent had evaporated leaving solid powder on the bottom of the
container. Not the C6 phase is still present.

C3/Sr2Nb3O10 in 2P (Sample 14c) SN Sediment

3 5 7 9 11 13
2θ (Deg)

Figure 7.16. XRD pattern for C3/Sr2Nb3O10 in 2P SN sediment (Sample 14c). Note the presence of
both C3 and hydrated phases at 4.8 and 5.7 o, respectively.

177
C6/Sr2Nb3O10 in 5M2H (Sample 15c) SN Sediment

2 4 6 8 10 12 14
2θ (Deg)

Figure 7.17. XRD pattern for C6/Sr2Nb3O10 in 5M2H (Sample 15c) Sediment. The stacking reflection
at 5.8o is consistent with a partially hydrated phase. The slight plateau between 3.5 and 4.7o is possibly
indicative of remaining C6/Sr2Nb3O10.

7.4.2 Dried Suspensions

The fundamental premise of this exfoliation is built upon the necessity for alkyl

grafting into the interlayer gallery in order for organic solvent to be employed, and it is

naturally important to determine the behavior of these groups upon exfoliation. After

sitting for over one week, several drops of suspension from Samples 4 and 5

(C6/Ca2Nb3O10 and C6/Sr2Nb3O10, respectively, in 5M2H) were air dried onto a glass plate

and used to obtain the XRD patterns shown in Figure 7.18. Both patterns show broad

stacking reflections due to the C6 phase consistent with either reduced particle size from

the exfoliation procedure and/or disordered restacking of nanosheets during drying.

Sample 4 also showed a small stacking reflection due to the hydrated phase. As subsequent

results will show, the point at which reversion to the hydrated phase and the degree to

which this occurs are both ambiguous and difficult to accurately determine experimentally

178
C6/Ca2Nb3O10 in 5M2H Dried Suspension Sample 4

2.5 4.5 6.5 8.5 10.5 12.5


2θ (Deg)

C6/Sr2Nb3O10 in 5M2H Dry Suspension Sample 5

2.5 4.5 6.5 8.5 10.5 12.5


2θ (Deg)

Figure 7.18. XRD pattern for C6/Ca 2Nb3O10 in 5M2H suspension (Sample 4) and C6/Sr2Nb3O10 in
5M2H suspensions (Sample 5) obtained by air-drying several drops of the suspension (before
centrifugation) onto a glass plate. The reflections at ~3.6 and ~5.8 degrees correspond to C6 grafted
and hydrated phases, respectively. Note the absence of a strong hydrated phase reflection in Sample
5 is possibly due to differences in the suspension preparation (sonication vs. microwave heating),
differences in the procedure employed for synthesis of the parent compound (conventional heating for
Sample 4 vs. microwave heating for Sample 5), or differing composition of the parent compound.

Based on the success of adding a microwave heating step, a sample was prepared

using a conventional heating step to serve as a comparison. Sample 7 was prepared by

mixing C6/Sr2Nb3O10 with 5M2H and heating in a steel bomb for 3 days at 150 oC.

Immediately after transferring the sample to a glass vial post-heating, it was observed that

along with some of the powder settling out of solution within minutes, there was a distinct

179
portion of suspension settling out over the course of hours (designated moderate settling

phase) after which the suspension remained stable (designated slow settling phase). Using

a pipette, several drops were taken from the fast settling and slow settling phases and dried

on glass slides for comparisons with XRD (see Figure 7.19). The fast settling phase

contained an intense reflection related to the C6 phase and a smaller reflection related to

the hydrated phase while the slow settling phase showed only a hydrated phase reflection.

This suggests a greater degree of exfoliation results in reversion to the hydrated phase.

Considering exfoliation leads to more accessible sheet surface sites, it is logical that grafted

alkyl groups would then be more easily removed. However, the presence of C6 grafted

phase in dried suspensions prepared using microwave heating, even after sitting untouched

for long periods of time, indicates that rapid heating reduces the susceptibility to alkyl

group loss. The fate of the grafted alkyl groups will be discussed further in section 7.7.

180
C6/Sr2Nb3O10 in 5M2H Sample 7 Slow Settling Phase

2 4 6 8 10 12 14
2θ (Deg)

C6/Sr2Nb3O10 in 5M2H Sample 7 Moderate Settling Phase

1.5 3.5 5.5 7.5 9.5 11.5 13.5


2θ (Deg)

Figure 7.19. XRD patterns for C6/Sr2Nb3O10 in 5M2H (Sample 7) from the slow settling phase (Top)
and moderate settling phase (Bottom).

As the results of the initial grafting work show (see Chapter 6), the longer alkyl

chains in C6 compounds create a more hydrophobic interlayer environment leading to

substantially less reversion to the hydrated phase over time when compared to C1 and C3

compounds. This trend is also seen with exfoliated samples. Figure 7.20 shows XRD

patterns of dried suspensions of C3/CaSrNb3O10 in 2P (Sample 9) prepared the same day

the suspension was prepared (Bottom pattern) and 2 days after sitting untouched (Top

pattern). While the same day pattern shows both a C3 phase and a hydrated phase, after 2

181
days the C3 stacking reflection has been reduced almost completely while the hydrated

stacking reflection has increased intensity. This further suggests C6 grafted compounds

are better suited for efficient exfoliation.

C3/CaSrNb3O10 in 2P Sample 9 Dried Suspension


(Two Weeks)

2.5 4.5 6.5 8.5 10.5 12.5


2θ (Deg)

C3/CaSrNb3O10 in 2P Sample 9 Dried Suspension


(Day of Synthesis)

2.5 4.5 6.5 8.5 10.5 12.5


2θ (Deg)

Figure 7.20. XRD pattern for C3/CaSrNb3O10 in 2P (Sample 9) dried suspension. The pattern on top
is from a dried suspension prepared two weeks after the original synthesis while the bottom pattern is
from the day of the synthesis. The lack of a grafted stacking reflection, in the dried suspension
prepared after sitting two weeks, and increased intensity of the hydrated stacking reflection suggest
sheets in the suspension are reverting from the grafted state to the hydrated state.

182
7.4.3 Solid From SN

While the lack of particle sedimentation in these colloidal suspensions after sitting

untouched over long periods of time is consistent with successful exfoliation into single

nanosheets, partially exfoliated sheets (still stacked, but to a lesser degree) can also remain

suspended if only the force of gravity is inducing sedimation. To ensure only fully

exfoliated sheets are present in solution, samples were centrifuged at 6000 rpm for 45

minutes. The centrifugal force of 4303 g under these conditions induces sedimentation of

partially exfoliated sheets while fully exfoliated sheets remain in the supernatant. Absence

of an observed Tyndall Effect in the supernatant of all samples except Sample 15c (a 2nd

exfoliation) confirmed that partially exfoliated sheets were present in the original

suspensions. To determine whether nanosheets were still present in the supernatant at all,

the solvent was removed and the remaining solid studied with XRD.

Four methods of solvent removal were tested: (1) Rapid evaporation by dropping

SN onto a glass plate heated to 80 oC; (2) Slow evaporation by dropping SN onto a glass

plate at room temperature; (3) Slow evaporation in a N2 atmosphere; and (4) freeze-drying

the supernatant such that solvent is sublimed away. As seen in Figure 7.21, rapid drying

of the SN resulted in the observance of broad low intensity reflections in the XRD pattern

associated with the hydrated phase. Simulation of the (001) stacking reflections as a

function of the number of sheets stacked together204 (shown in grey in Figure 7.21) matches

with the observed stacking reflection assuming an average number of 2 sheets are stacked

together. The simulated pattern was calculated using the Laue function:205

183
2𝜋𝑁𝑐𝑠𝑖𝑛𝜃
1+𝑐𝑜𝑠2 2𝜃 𝑠𝑖𝑛2 ( )
Equation 7.1: 𝐼00𝑙 (𝑡ℎ) = 2
|𝐹00𝑙 (𝜃)|2 𝜆
2𝜋𝑐𝑠𝑖𝑛𝜃
𝑠𝑖𝑛 𝜃𝑐𝑜𝑠𝜃 𝑁𝑠𝑖𝑛2 ( )
𝜆

where N is the number of stacked sheets and F00l is the structure factor, calculated using a

tetragonal Sr2Nb3O10 crystal structure. Reflections at 23.1o (3.85 Å), 28.7o (3.10 Å), and

32.7o (2.74 Å) correspond to lateral perovskite dimensions confirming the perovskite

structure remains intact throughout the exfoliation process. No discernable stacking

reflections associated with the C6 phase were observed.

C6/Sr2Nb3O10 in 5M2H SN Sample 5 Rapid Drying


23.1O

28.7O

32.7O Supernatant

Fit

Blank Glass Plate

0 450
2 6 10 14 18 22 26 30 34 38 42 46
2θ Deg

Figure 7.21. XRD pattern for C6/Sr2Nb3O10 in 5M2H centrifuge supernatant (Sample 5) after drying.
The sample was rapidly dried by placing drops of the supernatant onto a glass plate heated to 80 oC
such that solvent evaporated within seconds until a significant amount of solid remained. The broad
hump in the pattern is due to the background signal of the glass plate (blank plate shown in blue). The
grey pattern is a simulation of (00l) stacking reflections assuming an average number of 2 hydrated
sheets stacked together.

184
Figures 7.22 and 7.23 show XRD patterns for SNs of C3/Sr2Nb3O10 in 2P (Sample

8) slowly dried in air and C6/Sr2Nb3O10 in 5M2H (Sample 13) dried in a N2 atmosphere,

respectively. Compared to the rapid drying method, reflections in the XRD patterns of

these samples are much sharper. This indicates slow evaporation leads to more ordered

restacking of the exfoliated sheets, although they are still significantly more disordered

than the parent compounds. It was thought that drying in a N2 atmosphere would protect

sheets from reverting to the hydrated phase (assuming grafted alkyl groups are still present

in the SN), but again only the hydrated phase was observed.

C3/Sr2Nb3O10 in 2P Sample 8 SN
002

200 220
004

3 9 15 21 27 33 39 45 51 57
2θ (Deg)

C3/Sr2Nb3O10 in 2P Sample 8 SN
002

004

2 4 6 8 10 12 14
2θ (Deg)

Figure 7.22. Long-range (Top) and short-range (Bottom) XRD patterns for C3/Sr2Nb3O10 in 2P
(Sample 8) supernatant after sitting for 2 weeks. Only the hydrated phase is observed with fewer
higher order reflections compared to parent compound. Note the slight hump at ~3 o is due to the glass
slide being place in a position such that it blocked the X-Ray beam below 3o.

185
C6/Sr2Nb3O10 in 5M2H Sample 13 Dried SN

3 13 23 33 43 53
2θ (Deg)

Figure 7.23. Long-range XRD pattern of C6/Sr2Nb3O10 in 5M2H (Sample 13) supernatant dried in a
N2 atmosphere. The lack of many higher order reflections is consistent with irregular restacking of
sheets, though the stacking reflection position is consistent with a hydrated phase rather than the
desired grafted phase.

As discussed in section 7.2, freeze-drying of nanosheet suspensions has been used

in previous work to successfully extract nanosheets out of solution without restacking.

Figure 7.24 shows the resulting XRD pattern for the SN of C6/Sr2Nb3O10 in 5M2H (Sample

14), freeze-dried using t-butanol as a co-solvent. As will be discussed further in section

7.7, the freeze-drying process for these exfoliated systems presents inherent difficulties and

has not yet been optimized. As a result, at several points in the freeze-drying process, the

sample began to liquify and was refrozen. This is the likely cause of the restacking

observed in the XRD pattern. As with the other methods of solvent removal, only the

hydrated phase was observed in the final product.

186
C6/Sr2Nb3O10 Sample 14 Freeze-Dried Product

3 13 23 33 43 53
2θ (Deg)

Figure 7.24. Long-range XRD pattern for C6/Sr2Nb3O10 after freeze-drying in a 5M2H/t-BuOH
mixture (Sample 12). The presence of a sharp stacking reflection and higher order reflections is likely
due to the frozen solution melting several times during the freeze-drying process, essentially resulting
in a slow evaporation of solvent and thus allowed for ordered restacking of sheets.

7.5 SEM and EDS Results

Figure 7.25 shows an SEM image of C6/Sr2Nb3O10 in 5M2H (Sample 3) treated

with only sonication. The typical layered niobate parent compounds exist as thin platelets

with lateral dimensions of 1-3 um and platelet thickness of several hundred nanometers

(see chapter 4 for figures and discussion). After sonication, Figure 7.26 clearly shows that

particle size has been reduced dramatically such that the maximum particle size observed

became 681 and 154 nm in the lateral and stacking directions, respectively, while most

particles are too small and irregular to be clearly pictured and/or measured. This is

consistent with generation of a colloidal suspension. Furthermore, SEM data for the SN

of C6/Ca2Nb3O10 in 5M2H (Sample 4, Figure 7.26) showed that after centrifugation, only

irregularly shaped particles lacking the typical platelet shape with lateral dimensions less

than ~500 nm and a thickness too small to be measured with precision are observed. This

187
is consistent with irregular restacking of sheets observed in several other studies on niobate

nanosheets.182, 188, 198

Figure 7.25. SEM image of C6/Sr2Nb3O10 in 5M2H (Sample 3) on TESCAN instrument. While the
particles have the platelet shape seen in parent compounds, the particle size is substantially smaller
with 681 nm being the largest lateral dimension seen in the sample. Plate thickness of 67 and 154 nm
correspond to ~40 and ~100 sheets stacked together, respectively.

Figure 7.26. C6/Ca2Nb3O10 in 5M2H SN (Sample 4 sitting unmoved for three weeks, Left) and
C6/Sr2Nb3O10 in 5M2H SN (Sample 6, Right) SEM images. Sample was prepared by placing several
drops onto SEM sample holder and letting air dry. Sample 6 on the right was prepared over a much
longer period with ~1mL of SN in total dried onto the Si wafer. Note the substantially increased
irregularity and decreased size of the particles compared to parent compounds.

188
Of course, considering that even the slightest agitation after centrifugation can lead

to sedimentary particles entering the supernatant, combined with the irregular shape of

particles observed, it was necessary to use a more thorough procedure to prevent

contamination and confirm the particles were indeed niobates. The supernatant of

C6/Sr2Nb3O10 in 5M2H (Sample 6), after centrifuging for 45 min at 6000 rpm, was rapidly

dried under heat, followed by dispersing the solid residue in 10 mL of 5M2H and

centrifuging again for 45 min at 6000 rpm. This final supernatant was placed dropwise

onto a Si wafer for SEM analysis (see Figure 7.27). The presence of Nb in the EDS

spectrum combined with a stacking reflection seen in the XRD pattern (Figure 7.21) for

the same sample is strong evidence that exfoliated nanosheets were successfully produced.

cps/eV
C6Sr2Nb3O10 5-methyl-2-hexanone supernatant 2

4
Nb
C O Na Si Nb

0
1 2 3 4 5 6
keV

Figure 7.27. SEM image (top) and EDS spectrum (bottom) for C6/Sr2Nb3O10 in 5M2H (Sample 5)
supernatant after evaporated solvent, re-dissolving, centrifuging, and final evaporation of solvent. The
small and irregular shape of the particles combined with the presence of Nb in the EDS spectrum is
strong evidence for successful exfoliation of nanosheets. Note that there is no Sr signal assigned in the
EDS spectrum due to the Si signal from the wafer overlapping.

As Figure 7.27 above shows, while Nb is observed in the EDS spectrum, Sr cannot

be observed due to the large Si signal. To overcome this, a new sample preparation method

was employed. A piece of double sided copper tape was placed on an aluminum SEM

189
sample holder. Acetone was used to strip away the adhesive on the copper, due to the

presence of Si in the adhesive, and the nanosheet suspension was air dried dropwise onto

the copper at room temp (~1mL added in total. As was expected, the resulting SEM images

contained irregularly shaped particles <1 micrometer in length, consistent with previous

images and the corresponding EDS spectra contain both Nb and Sr signals with the desired

compositional ratios (when error is considered – see Table 7.4).

a b
Figure 7.28. SEM images for (a) C3/Sr2Nb3O10 in 2P (sample 14c) SN, (b) C6/Sr2Nb3O10 in 5M2H
(sample 15c) SN. Note the horizontal lines in image (a) are to due scratches on the copper tape
background created during removal of the adhesive layer.

190
cps/eV
8 C6Sr2 5M2H 106a 4

Nb
C O Cu Sr Nb
4

0
0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
keV

Figure 7.29. EDS spectrum of C6/Sr2Nb3O10 in 5M2H SN (Sample 15c).

Table 7.4. Atomic composition of samples 14C and 15C using copper tape as substrate. The letter in
parentheses indicates to which image in Figure 7.28 above the data correspond. Error was
automatically calculated by the Quantax EDS application.
Sample Nb% error Sr% error
14c (a) 0.96 0.2 0.73 0.1
15c (b) 2.97 0.9 2.01 0.5

In an effort to overcome SEM resolution problems when imaging nanosheets as

well as expand the potential applications of these materials, sample 15c SN was treated

with a suspension of Cu2O nanocubes in order to deposit the nanosheets on the cube

surface. Approximately 1 mL of both nanosheet SN (C6/Sr2Nb3O10 in 5M2H) and

nanocube suspension were mixed together vigorously. One drop of this mixture was placed

on a Si wafer and left to air dry, followed by gold coating and SEM imaging. SEM images

(Figures 7.30 and 7.31) show the presence of nanosheets on the Cu2O nanocube surface.

The image in 7.30 was taken with a lower beam voltage leading to decreased resolution,

191
but due to the tendency for the electron beam to pass through sheets at higher power,

significant aggregation of sheets was observed with lateral dimensions less than 500 nm

and irregular shapes. Figure 7.31 shows an image using higher beam voltage and thus

improved resolution, although the sheets have an almost transparent appearance due to the

aforementioned effect of thin particles.

Figure 7.30. SEM image of C6/Sr2Nb3O10 nanosheets deposited onto a Cu2O nanocube. The small (less
than 500 nm), irregularly-shaped particles are the exfoliated/restacked nanosheets. Note that there is
a wide distribution of nanosheet agglomeration.

192
Figure 7.31. SEM image of C6/Sr2Nb3O10 nanosheets deposited onto a Cu2O nanocube. The small (less
than 500 nm), irregularly-shaped particles are the exfoliated/restacked nanosheets. Note that the
almost transparent quality of nanosheets on the cube surface is consistent with the electron beam
passing through the thin sheet.

7.6 TGA Results and Exfoliation Efficiency

To quantify the degree of exfoliation, both 5M2H and 2P samples, along with a

suspension produced using the conventional TBAOH exfoliation, underwent Thermal

Gravimetric Analysis. By measuring the initial mass of a droplet of each sample along

with the final mass after heating to 200 oC (along with an isothermal stage to ensure all

solvent had been evaporated) the w/v concentration in the original solution was calculated.

Using that concentration and the theoretical maximum concentration assuming 100%

exfoliation it was found the 5M2H sample underwent 25% exfoliation, the 2-pentanone

193
sample underwent 28% exfoliation, and the conventional TBAOH sample underwent 22%

exfoliation.

7.7 Discussion

7.7.1 Role of Solvent and Parent Compound

While both C3 and C6 parent compounds were successfully exfoliated, the

tendency for C3 compounds to revert to a hydrated state at a much faster rate suggests that

the C6 phase is better suited for successive exfoliations. As the TGA data show, initial

exfoliation at best leads to ~25% of the sample being successfully exfoliated. Thus

successive exfoliations have clear benefits. While the microwave grafting technique

developed in this work was not successful for long chain alcohols beyond 1-hexanol, longer

chain alcohols would likely lead to greater interlayer hydrophobicity and thus further

prevent hydration.

Choice of solvent for exfoliation is less clear and more dependent on the goals of

the exfoliation. While any solvent used would need to be a relatively close HSP match,

the results of the same C6 parent compound using either 2P or 5M2H (samples 10 and 11,

respectively) show that increased microwave susceptibility can overcome deviations from

ideal HSP values in terms of fraction of sample exfoliated. However, the suspension color

generated after microwave heating was substantially more dramatic in 5M2H for all

samples tested. While the origins of this change need further elucidation, 5M2H is superior

to 2P should generation of a colored suspension be proven desirable. Testing a wider

variety of solvents is undoubtedly a desirable route for future research into this technique.

194
7.7.2 Role of Sonication

Sonication was found to be very useful in the exfoliation procedure. The initial

sonication step, in and of itself, produces colloidal suspensions consistent with exfoliation

and is important to improved exfoliation efficiency. This is particularly true of samples

where the solvent is a close HSP match to the grafted parent compound. The importance

of initial sonication for samples without a close HSP match is more ambiguous and needs

further study to determine its necessity. Post-heating sonication has not been shown to

significantly affect exfoliation behavior for better or for worse. Final sonication was

included in the optimized procedure due to its simplicity, relatively short time, and the little

attention required by the operator during this step.

7.7.3 Role of Microwave Heating

As with the grafting procedure, microwave heating was found to be key to efficient

exfoliation. Exfoliated suspensions consistently developed noticeably more opaqueness

post-heating as well as producing more opaque suspensions when heated without initial

sonication than with sonication alone. Furthermore, only samples with a heating stage

produce colored suspensions. Microwave heating is also preferable to conventional

heating not only due to its substantially more rapid nature, but also because conventional

heating resulted in more susceptibility to hydration thus inhibiting the effectiveness of

repeated exfoliations. The stability of the perovskite structure after undergoing such

195
extreme conditions of high temperature and pressure highlight the broad range of systems

in which these nanosheets can be used.

7.7.4 Treatment of Supernatant

The supernatant treatment methods used in this work including air-drying, drying

under heat, drying in a N2 atmosphere, and freeze-drying have been useful in obtaining

indirect experimental confirmation of successful nanosheet exfoliation in solution.

However, collecting sheets out of solution using these techniques has inevitably led to

restacking, albeit in a disordered fashion. Potential applications of this exfoliation

technique are by no means limited to nanosheets out of solution, but the ability to collect

sheets out of solution without restacking would certainly increase the technique’s potential.

As has been discussed previously, freeze-drying is the most widely used method of

collecting exfoliated samples, but the freezing points of the solvents used in this work are

too low to be used in our freeze-drying set up. Indeed, most freeze-drying systems are not

equipped to handle such compounds, but the use of a co-solvent has been applied with

much success to systems with similar limitations.

7.7.5 Fate of the Grafted Alkyl Groups

One effect of exfoliation of layered niobates into single nanosheets is dramatically

increased access to the perovskite surface sites. This is particularly useful in their use as

solid acid catalysts due to the highly acidic protonated environment on the surface of the

sheets. In these grafted materials, however, this means alkyl groups are more easily

196
accessed for potential removal by hydration. This is not inherently disadvantageous as

alkyl group loss produces new protonated environments that can be accessed, but if alkyl

group loss is unavoidable then it is a limitation to the controlled design of nanosheets in

the sense that initially grafted groups can be further reacted. However, it is possible that

observations of alkyl group loss in this work are the result of unoptimized reaction

conditions or collection procedures.

Evidence of alkyl group loss and subsequent generation of the protonated site in

this work is seen most clearly in XRD patterns. As shown in section 7.4, each successive

stage in the exfoliation process from the parent compound to the initial colloidal suspension

to the final centrifuged SN shows progressive loss of the grafted stacking reflection and

gain of the hydrated phase stacking reflection. However, it is not evident whether alkyl

group loss occurs during the exfoliation itself or whether it is the result of solvent removal.

While 1H and 13C NMR have been used to distinguish between grafted alkyl groups and

nongrafted alcohols by Sugahara et al.,64-66, 68 the dilute nature of the suspensions would

make observation of the relevant signals difficult. Furthermore, if a successful freeze-

drying method can be implemented, it may be possible to observe alkyl groups on extracted

nanosheets using the same methods used in chapter 6.

There also exists the possibility that alkyl groups are still present after solvent

removal, but are unobserved in subsequent XRD patterns. The flexibility of the alkyl

groups within the interlayer can be most clearly illustrated by the range of positions where

the C6 phase stacking reflections are observed (See Table 7.2). Should grafted sheets be

restacking in excessive disorder, a clear stacking reflection may not be readily observable

197
in XRD patterns. A test of this hypothesis requires an attempted 1-hexanol grafting of the

final product post solvent removal.

7.8 Conclusions

Using rapid microwave heating generating high temperatures and pressures, a new

method of nanosheet exfoliation from grafted parent compounds in organic solvents has

been successfully developed. Visual observations of colloidal suspensions not settling

under gravitational force is strongly indicative of the presence of small exfoliated (or

partially exfoliated at this point) particles and XRD patterns confirm the presence of the

parent layered niobates. Centrifugation at 6000 rpm for 45 minutes served to remove any

partially exfoliated sheets from solution. After removing the exfoliation solvent from the

SN, SEM images of irregularly shaped particles, the presence of Nb in EDS spectra, and

highly disordered stacking reflections in XRD patterns, along with observed Tyndall

effects in centrifuged suspensions, illustrate that layered niobates have been successfully

exfoliated into nanosheets. Furthermore, generation of a colored suspension serves as

inspiration for future studies into the behavior of nanosheets in organic suspensions.

The optimal exfoliation procedure found in this work consisted of an initial

sonication stage, a microwave heating stage using an acid digestion bomb, followed by a

final sonication stage. Varying combinations of solvent and parent grafted compound show

this method has the potential to be applied to a broad range of systems. Not only do TGA

data indicate the efficiency of this method is comparable to the conventional exfoliation

method using TBAOH, but also that unexfoliated material after initial treatment has been

198
primed for exfoliation in subsequent treatments. The rapid nature of this exfoliation

method, use of organic solvents, and exfoliation efficiency show it to be a promising

alternative to current exfoliation methods that broaden the range of systems niobate

nanosheets in which niobate nanosheets can be applied.

199
Chapter 8: Summary and Future Directions

The work detailed in this dissertation has been a synergistic combination of the

synthesis and the characterization of D-J Niobates. The compounds studied include

Rb(H)CaxSr2-xNb3O10, Rb(H)BaxSr2-xNb3O10, and (C1/C3/C6)/CaxSr2-xNb3O10 for x = 0-2,

and RbCa2NaNb4-xTaxO13 for x = 0 and 1.2. The synthetic techniques employed include

rapid microwave heating, MSS, acid exchanging, alcohol grafting (both conventional the

newly developed rapid microwave assisted method), and exfoliation (both using TBAOH

and the newly developed organic solvent method). Samples were characterized using

XRD, SEM, EDS, IR, UV-visible, and both liquid- and solid-state NMR (for 1H, 23Na, and
93
Nb nuclei). Using these synthetic and characterization methods, new and invaluable

insights into the structure and behavior of D-J Niobates have been elucidated. Should only

the most central results of this work be instilled upon the reader, they are as follows:
93
1. Nb NMR has exhibited heterogeneous cation distribution at the interstitial A-site

in RbCaxSr2-xNb3O10 and RbBaxSr2-xNb3O10 compounds. This contrasts the

observed homogeneity in the corresponding XRD patterns (both collected in this

work and in studies elsewhere). In fact, while many bulk/long range properties of

these compounds are consistent with homogeneous cation distribution, the local

structures observed by NMR show a clear heterogeneity. This dichotomy of cation

distribution not only explains the grafting results presented in this work, but

provides insight into the potential intelligent design of future perovskite materials.
93 23
2. Nb and Na NMR was employed to determine cation site ordering within the

four-layer Rb(H)Ca2NaNb4-xTaxO13 structure. Despite the multiple overlapping

200
signals in static NMR spectra, individual site parameters could be obtained using a

combination of MQMAS, MQMAS-t2-REDOR, and BRAIN-CP experiments as

well as collecting static spectra at multiple magnetic field strengths. It was

ultimately determined that Ta has a strong site preference for interior octahedra.

The increased ionic character of Ta-O bonds compared to Nb-O bonds reduces the

ability to undergo second-order Jahn-Teller distortions and thus it is expected Ta

would prefer the more symmetrical interior octahedral site.

3. A rapid microwave-assisted method of grafting n-alcohols up to hexanol into the

perovskite interlayer gallery was developed that reduced synthesis time by 97%

when compared to the conventional heating method without substantial loss of alkyl

group coverage. It was found that while HCa2Nb3O10 does not undergo rapid

grafting beyond the methylated stage, substitution of as little as 10% Sr into the A-

site imbues compounds with grafting ability equal to that of the pure HSr2Nb3O10

form. The heterogeneous cation distribution observed in 93Nb NMR of mixed Ca/Sr

compounds explains this trend in grafting ability and shows that grafting ability can

be imbued without substantially altering other bulk properties.

4. A new organic solvent based nanosheet exfoliation method was successfully

developed using grafted niobates. The expanded interlayer gallery reduces

electrostatic forces holding sheets together and a combination of sonication and

microwave heating in an acid digestion bomb introduces organic solvent into the

interlayer gallery such that the dispersion forces of the alkyl groups are weakened

and the niobate sheets break apart until only single sheets remain. A close HSP

201
match between the grafted alcohol and organic solvent (accounting for the reduced

hydrogen bonding interactions of the alcohol post-grafting) is a strong indicator of

the solvent’s exfoliation ability, though high microwave susceptibility can

overcome a slight HSP mismatch. A fascinating, but still not conclusively

explained, phenomenon of this exfoliation procedure was the generation of strongly

yellow colored suspensions. Further study is required to properly identify the origin

of these colored suspensions, but preliminary data in this work suggest it is likely

a charge-transfer complex between the ketone and nanosheet surface is the culprit.

While I may not be personally involved in future research directions, there remains an

abundance of questions left to answer:

• What is the nature of the color generated after exfoliation?

• Why does grafting effect the band gap and can it be controlled?

• How can exfoliated sheets be collected out of solution?

• What are the catalytic properties of these materials before and after exfoliation?

• What other groups can be grafted into the interlayer and can they be modified

post-exfoliation without being removed from the surface?

• How can 93Nb NMR be used in colloidal solutions?

• Will polarization transfer-based NMR experiments help characterize H-O-Nb

environment and elucidate grafting ability?

• Would it be possible to design a high-temperature and high-pressure sample

rotor to monitor grafting/exfoliation in situ?

202
• Can a combination of microwave heating and conventional heating be used to

graft HCa2Nb3O10 beyond the C1 stage?

Despite the new insights into layered niobate structure and properties this work has

produced, it is but a mere glimpse of the vast unknown still hidden in the darkness, waiting

for inevitable illumination.

203
References

1. Dimos, D.; Mueller, C., Perovskite Thin Films for High-Frequency Capacitor
Applications 1. Annual Review of Materials Science 1998, 28 (1), 397-419.
2. Izuha, M.; Abe, K.; Fukushima, N., Electrical properties of all-perovskite oxide
(SrRuO3/BaxSr1-xTiO3/SrRuO3) capacitors. Japanese journal of applied physics
1997, 36 (9S), 5866.
3. Kim, Y.-H.; Kim, H.-J.; Osada, M.; Li, B.-W.; Ebina, Y.; Sasaki, T., 2D Perovskite
Nanosheets with Thermally-Stable High-κ Response: A New Platform for High-
Temperature Capacitors. ACS applied materials & interfaces 2014, 6 (22), 19510-
19514.
4. Ahn, C.-W.; Choi, J.-J.; Ryu, J.; Yoon, W.-H.; Hahn, B.-D.; Kim, J.-W.; Choi, J.-
H.; Park, D.-S., Composition design rule for energy harvesting devices in
piezoelectric perovskite ceramics. Materials Letters 2015, 141, 323-326.
5. Kim, Y.-J.; Dang, T.-V.; Choi, H.-J.; Park, B.-J.; Eom, J.-H.; Song, H.-A.; Seol, D.;
Kim, Y.; Shin, S.-H.; Nah, J., Piezoelectric properties of CH 3 NH 3 PbI 3
perovskite thin films and their applications in piezoelectric generators. Journal of
Materials Chemistry A 2016, 4 (3), 756-763.
6. Cheng, L.-Q.; Wang, K.; Li, J.-F., Synthesis of highly piezoelectric lead-free (K,
Na) NbO3 one-dimensional perovskite nanostructures. Chem. Commun. 2013, 49
(38), 4003-4005.
7. Paul, A. K.; Reehuis, M.; Ksenofontov, V.; Yan, B.; Hoser, A.; Többens, D. M.;
Abdala, P. M.; Adler, P.; Jansen, M.; Felser, C., Lattice Instability and Competing
Spin Structures in the Double Perovskite Insulator Sr 2 FeOsO 6. Physical Review
Letters 2013, 111 (16), 167205.
8. Torrance, J.; Lacorre, P.; Nazzal, A.; Ansaldo, E.; Niedermayer, C., Systematic
study of insulator-metal transitions in perovskites R NiO 3 (R= Pr, Nd, Sm, Eu) due
to closing of charge-transfer gap. Physical Review B 1992, 45 (14), 8209.
9. De Teresa, J.; Ibarra, M.; Garcia, J.; Blasco, J.; Ritter, C.; Algarabel, P.; Marquina,
C.; Del Moral, A., Spin-Glass Insulator State in (Tb-La) 2/3 C a 1/3 Mn O 3
Perovskite. Physical Review Letters 1996, 76 (18), 3392.
10. Michel, C.; Er-Rakho, L.; Raveau, B., The oxygen defect perovskite
BaLa4Cu5O13. 4, a metallic conductor. Materials research bulletin 1985, 20 (6),
667-671.
11. Mizusaki, J., Nonstoichiometry, diffusion, and electrical properties of perovskite-
type oxide electrode materials. Solid State Ionics 1992, 52 (1), 79-91.
12. Xu, X.; Randorn, C.; Efstathiou, P.; Irvine, J. T., A red metallic oxide
photocatalyst. Nature materials 2012, 11 (7), 595-598.
13. Jung, J. I.; Jeong, H. Y.; Lee, J. S.; Kim, M. G.; Cho, J., A bifunctional perovskite
catalyst for oxygen reduction and evolution. Angewandte Chemie 2014, 126 (18),
4670-4674.

204
14. Jin, C.; Yang, Z.; Cao, X.; Lu, F.; Yang, R., A novel bifunctional catalyst of Ba 0.9
Co 0.5 Fe 0.4 Nb 0.1 O 3− δ perovskite for lithium–air battery. International
Journal of Hydrogen Energy 2014, 39 (6), 2526-2530.
15. Lin, K.-H.; Wang, C.-B.; Chien, S.-H., Catalytic performance of steam reforming
of ethanol at low temperature over LaNiO 3 perovskite. International Journal of
Hydrogen Energy 2013, 38 (8), 3226-3232.
16. Rubel, M. H.; Miura, A.; Takei, T.; Kumada, N.; Mozahar Ali, M.; Nagao, M.;
Watauchi, S.; Tanaka, I.; Oka, K.; Azuma, M., Superconducting Double Perovskite
Bismuth Oxide Prepared by a Low‐Temperature Hydrothermal Reaction.
Angewandte Chemie International Edition 2014, 53 (14), 3599-3603.
17. Schoop, L. M.; Müchler, L.; Felser, C.; Cava, R., Lone pair effect, structural
distortions, and potential for superconductivity in Tl perovskites. Inorganic
chemistry 2013, 52 (9), 5479-5483.
18. Lee, M. M.; Teuscher, J.; Miyasaka, T.; Murakami, T. N.; Snaith, H. J., Efficient
Hybrid Solar Cells Based on Meso-Superstructured Organometal Halide
Perovskites. Science 2012, 338 (6107), 643-647.
19. Kim, H.-S.; Lee, C.-R.; Im, J.-H.; Lee, K.-B.; Moehl, T.; Marchioro, A.; Moon, S.-
J.; Humphry-Baker, R.; Yum, J.-H.; Moser, J. E.; Grätzel, M.; Park, N.-G., Lead
Iodide Perovskite Sensitized All-Solid-State Submicron Thin Film Mesoscopic
Solar Cell with Efficiency Exceeding 9%. Scientific Reports 2012, 2, 591.
20. Goldschmidt, V., Crystal structure and chemical constitution. Transactions of the
Faraday Society 1929, 25, 253-283.
21. Momma, K.; Izumi, F., VESTA 3 for three-dimensional visualization of crystal,
volumetric and morphology data. Journal of Applied Crystallography 2011, 44 (6),
1272-1276.
22. Brown, I. D., The Chemical Bond in Inorganic Chemistry: The Bond Valence
Model. Oxford University Press: New York, 2002; p 278.
23. Dion, M.; Ganne, M.; Tournoux, M., Nouvelles familles de phases
MIMII2Nb3O10 a feuillets “perovskites”. Materials research bulletin 1981, 16
(11), 1429-1435.
24. Liang, Z.-H.; Tang, K.-B.; Chen, Q.-W.; Zheng, H.-G., RbCa2Nb3O10 from X-ray
powder data. Acta Crystallographica Section E 2009, 65 (6), i44.
25. Toda, K.; Sato, M., Synthesis and structure determination of new layered
perovskite compounds, ALaTa2O7 and ACa2Ta3O10 (A= Rb, Li). J. Mater.
Chem. 1996, 6 (6), 1067-1071.
26. Arney, D.; Porter, B.; Greve, B.; Maggard, P. A., New molten-salt synthesis and
photocatalytic properties of La 2 Ti 2 O 7 particles. Journal of Photochemistry and
Photobiology A: Chemistry 2008, 199 (2), 230-235.
27. Li, L.; Deng, J.; Chen, J.; Sun, X.; Yu, R.; Liu, G.; Xing, X., Wire structure and
morphology transformation of niobium oxide and niobates by molten salt synthesis.
Chemistry of Materials 2009, 21 (7), 1207-1213.
28. Porob, D. G.; Maggard, P. A., Flux syntheses of La-doped NaTaO 3 and its
photocatalytic activity. Journal of Solid State Chemistry 2006, 179 (6), 1727-1732.

205
29. Geselbracht, M. J.; Walton, R. I.; Cowell, E. S.; Millange, F.; O'Hare, D., An
Investigation of the Synthesis of the Layered Perovskite RbCa2Nb3O10 Using
Time-Resolved in Situ High-Temperature Powder X-ray Diffraction. Chemistry of
Materials 2002, 14 (10), 4343-4349.
30. Thangadurai, V.; Schmid-Beurmann, P.; Weppner, W., Synthesis, Structure, and
Electrical Conductivity of A′[A2B3O10](A′=Rb, Cs; A=Sr, Ba; B=Nb, Ta): New
Members of Dion–Jacobson-Type Layered Perovskites. Journal of Solid State
Chemistry 2001, 158 (2), 279-289.
31. Geselbracht, M. J.; White, H. K.; Blaine, J. M.; Diaz, M. J.; Hubbs, J. L.; Adelstein,
N.; Kurzman, J. A., New solid acids in the triple-layer Dion–Jacobson layered
perovskite family. Materials research bulletin 2011, 46 (3), 398-406.
32. Wang, X.; Adhikari, J.; Smith, L. J., An Investigation of Distortions of the Dion−
Jacobson Phase RbSr2Nb3O10 and Its Acid-Exchanged Form with 93Nb Solid
State NMR and DFT Calculations. The Journal of Physical Chemistry C 2009, 113
(40), 17548-17559.
33. Shannon, R. t.; Prewitt, C. T., Effective ionic radii in oxides and fluorides. Acta
Crystallographica Section B: Structural Crystallography and Crystal Chemistry
1969, 25 (5), 925-946.
34. Jacobson, A. J.; Johnson, J. W.; Lewandowski, J. T., Interlayer chemistry between
thick transition-metal oxide layers: synthesis and intercalation reactions of
K[Ca2Nan-3NbnO3n+1] (3 .ltoreq. n .ltoreq. 7). Inorganic chemistry 1985, 24 (23),
3727-3729.
35. Sato, M.; Kono, Y.; Jin, T., Structural Characterization and Ion Conductivity of
MCa2NaNb4O13 (M=Rb, Na) with Four Units of Perovskite Layer. Journal of the
Ceramic Society of Japan 1993, 101 (9), 980-984.
36. Maeda, K.; Mallouk, T. E., Comparison of two-and three-layer restacked Dion–
Jacobson phase niobate nanosheets as catalysts for photochemical hydrogen
evolution. Journal of Materials Chemistry 2009, 19 (27), 4813-4818.
37. Hata, H.; Kobayashi, Y.; Bojan, V.; Youngblood, W. J.; Mallouk, T. E., Direct
deposition of trivalent rhodium hydroxide nanoparticles onto a semiconducting
layered calcium niobate for photocatalytic hydrogen evolution. Nano letters 2008,
8 (3), 794-799.
38. Ebina, Y.; Sasaki, T.; Harada, M.; Watanabe, M., Restacked perovskite nanosheets
and their Pt-loaded materials as photocatalysts. Chemistry of materials 2002, 14
(10), 4390-4395.
39. Compton, O. C.; Carroll, E. C.; Kim, J. Y.; Larsen, D. S.; Osterloh, F. E., Calcium
niobate semiconductor nanosheets as catalysts for photochemical hydrogen
evolution from water. The Journal of Physical Chemistry C 2007, 111 (40), 14589-
14592.
40. Schaak, R. E.; Mallouk, T. E., Perovskites by design: a toolbox of solid-state
reactions. Chemistry of Materials 2002, 14 (4), 1455-1471.
41. Sasaki, T., Fabrication of nanostructured functional materials using exfoliated
nanosheets as a building block. Journal of the Ceramic Society of Japan 2007, 115
(1337), 9-16.

206
42. Ozawa, T. C.; Onoda, M.; Iyi, N.; Ebina, Y.; Sasaki, T., Bulk Functional Materials
Design Using Oxide Nanosheets as Building Blocks: A New Upconversion
Material Fabricated by Flocculation of Ca2Nb3O10– Nanosheets with Rare-Earth
Ions. The Journal of Physical Chemistry C 2014, 118 (3), 1729-1738.
43. Bao-Wen, L.; Minoru, O.; Kosho, A.; Yasuo, E.; Tadashi, C. O.; Takayoshi, S.,
Solution-Based Fabrication of Perovskite Multilayers and Superlattices Using
Nanosheet Process. Japanese journal of applied physics 2011, 50 (9S2), 09NA10.
44. Li, C.; Li, Y.; Xing, Y.; Zhang, Z.; Zhang, X.; Li, Z.; Shi, Y.; Ma, T.; Ma, R.;
Wang, K.; Wei, J., Perovskite Solar Cell Using a Two-Dimensional Titania
Nanosheet Thin Film as the Compact Layer. ACS applied materials & interfaces
2015, 7 (28), 15117-15122.
45. Hu, X.; Zhang, X.; Liang, L.; Bao, J.; Li, S.; Yang, W.; Xie, Y., High-Performance
Flexible Broadband Photodetector Based on Organolead Halide Perovskite.
Advanced Functional Materials 2014, 24 (46), 7373-7380.
46. Rao, C. N. R.; Ramakrishna Matte, H. S. S.; Maitra, U., Graphene Analogues of
Inorganic Layered Materials. Angewandte Chemie International Edition 2013, 52
(50), 13162-13185.
47. Zhuang, X.; Mai, Y.; Wu, D.; Zhang, F.; Feng, X., Two-Dimensional Soft
Nanomaterials: A Fascinating World of Materials. Advanced Materials 2015, 27
(3), 403-427.
48. Osada, M.; Takanashi, G.; Li, B.-W.; Akatsuka, K.; Ebina, Y.; Ono, K.; Funakubo,
H.; Takada, K.; Sasaki, T., Controlled Polarizability of One-Nanometer-Thick
Oxide Nanosheets for Tailored, High-κ Nanodielectrics. Advanced Functional
Materials 2011, 21 (18), 3482-3487.
49. Xu, T. G.; Zhang, C.; Shao, X.; Wu, K.; Zhu, Y. F., Monomolecular-Layer
Ba5Ta4O15 Nanosheets: Synthesis and Investigation of Photocatalytic Properties.
Advanced Functional Materials 2006, 16 (12), 1599-1607.
50. Tagusagawa, C.; Takagaki, A.; Hayashi, S.; Domen, K., Efficient utilization of
nanospace of layered transition metal oxide HNbMoO6 as a strong, water-tolerant
solid acid catalyst. Journal of the American Chemical Society 2008, 130 (23),
7230-7231.
51. Kitano, M.; Nakajima, K.; Kondo, J. N.; Hayashi, S.; Hara, M., Protonated titanate
nanotubes as solid acid catalyst. Journal of the American Chemical Society 2010,
132 (19), 6622-6623.
52. Takagaki, A.; Sugisawa, M.; Lu, D.; Kondo, J. N.; Hara, M.; Domen, K.; Hayashi,
S., Exfoliated nanosheets as a new strong solid acid catalyst. Journal of the
American Chemical Society 2003, 125 (18), 5479-5485.
53. Han, S.-T.; Zhou, Y.; Roy, V. A. L., Towards the Development of Flexible Non-
Volatile Memories. Advanced Materials 2013, 25 (38), 5425-5449.
54. Ma, R.; Sasaki, T., Nanosheets of Oxides and Hydroxides: Ultimate 2D Charge-
Bearing Functional Crystallites. Advanced Materials 2010, 22 (45), 5082-5104.
55. Tagusagawa, C.; Takagaki, A.; Hayashi, S.; Domen, K., Characterization of
HNbWO6 and HTaWO6 Metal Oxide Nanosheet Aggregates As Solid Acid
Catalysts. Journal of Physical Chemistry C 2009, 113 (18), 7.

207
56. Takagaki, A.; Yoshida, T.; Lu, D.; Kondo, J. N.; Hara, M.; Domen, K.; Hayashi, S.,
Titanium Niobate and Titanium Tantalate Nanosheets as Strong Solid Acid
Catalysts. Journal of Physical Chemistry B 2004, 108 (31), 7.
57. Kudo, A.; Kato, H.; Nakagawa, S., Water Splitting into H2 and O2 on New
Sr2M2O7 (M = Nb and Ta) Photocatalysts with Layered Perovskite Structures: 
Factors Affecting the Photocatalytic Activity. The Journal of Physical Chemistry B
1999, 104 (3), 571-575.
58. Shannon, R. D., Revised Effective Ionic Radii and Systematic Studies of
Interatomic Distances in Halids and Chalcogenides. Acta Crystallographica 1976,
A32, 751.
59. Huheey, J. E., Inorganic Chemistry. Harper and Row: New York, 1983.
60. Wen-Long;, Z.; Qing-Li;, Z.; Jin-Yun;, G.; Wen-Peng;, L.; Li-Hua;, D.; Shao-Tang,
R., Structures and luminescence properties of Yb3+ in the double perovskites
Ba2YB;O6 (B'=Ta5+, Nb5+) phosphors. Chin. Phys. B 2011, 20 (1), 016101.
61. Hong, Y.-S.; Kim, S.-J.; Kim, S.-J.; Choy, J.-H., B-site cation arrangement and
crystal structure of layered perovskite compounds CsLnTiNbO (Ln = La, Pr, Nd,
Sm) and CsCaLaTiNbO. Journal of Materials Chemistry 2000, 10 (5), 1209-1214.
62. Kim, H. J.; Byeon, S.-H.; Yun, H., Raman Spectra of the Solid-Solution between
Rb~ 2La~ 2Ti~ 3O~ 1~ 0 and RbCa~ 2Nb~ 3O~ 1~ 0. BULLETIN-KOREAN
CHEMICAL SOCIETY 2001, 22 (3), 298-302.
63. Maeda, K.; Mallouk, T. E., Comparison of two- and three-layer restacked Dion-
Jacobson phase niobate nanosheets as catalysts for photochemical hydrogen
evolution. Journal of Materials Chemistry 2009, 19, 4813-4818.
64. Takahashi, S.; Nakato, T.; Hayashi, S.; Sugahara, Y.; Kuroda, K., Formation of
Methoxy-Modified Interlayer Surface via the Reaction between Methanol and
Layered Perovskite HLaNb2O7.cntdot.xH2O. Inorganic Chemistry 1995, 34 (20),
5065-5069.
65. Suzuki, H.; Notsu, K.; Takeda, Y.; Sugimoto, W.; Sugahara, Y., Reactions of
Alkoxyl Derivatives of a Layered Perovskite with Alcohols:  Substitution Reactions
on the Interlayer Surface of a Layered Perovskite. Chemistry of Materials 2003, 15
(3), 636-641.
66. Tahara, S.; Sugahara, Y., Interlayer Surface Modification of the Protonated Triple-
Layered Perovskite HCa2Nb3O10·xH2O with n-Alcohols. Langmuir 2003, 19
(22), 9473-9478.
67. Tahara, S.; Takeda, Y.; Sugahara, Y., Preparation of Organic−Inorganic Hybrids
Possessing Nanosheets with Perovskite-Related Structures via Exfoliation during a
Sol−Gel Process. Chemistry of Materials 2005, 17 (24), 6198-6204.
68. Tahara, S.; Ichikawa, T.; Kajiwara, G.; Sugahara, Y., Reactivity of the Ruddlesden-
Popper Phase H2La2Ti3O10 with Organic Compounds: Intercalation and Grafting
Reactions. Chemistry of materials 2007, 19 (9), 2352-2358.
69. Shimada, A.; Yoneyama, Y.; Tahara, S.; Mutin, P. H.; Sugahara, Y., Interlayer
surface modification of the protonated ion-exchangeable layered perovskite
HLaNb2O7•xH2O with organophosphonic acids. Chemistry of Materials 2009, 21
(18), 4155-4162.

208
70. Bloch, F.; Hansen, W. W.; Packard, M., Nuclear Induction. Physical Review 1946,
69 (3-4), 127-127.
71. Purcell, E. M.; Torrey, H. C.; Pound, R. V., Resonance Absorption by Nuclear
Magnetic Moments in a Solid. Physical Review 1946, 69 (1-2), 37-38.
72. Ernst, R.; Anderson, W., Application of Fourier transform spectroscopy to
magnetic resonance. Review of Scientific Instruments 1966, 37 (1), 93-102.
73. Lauterbur, P. C., Image Formation by Induced Local Interactions: Examples
Employing Nuclear Magnetic Resonance. Nature 1973, 242 (5394), 190-191.
74. Stehling, M.; Turner, R.; Mansfield, P., Echo-planar imaging: magnetic resonance
imaging in a fraction of a second. Science 1991, 254 (5028), 43-50.
75. Laws, D. D.; Bitter, H.-M. L.; Jerschow, A., Solid-State NMR Spectroscopic
Methods in Chemistry. Angewandte Chemie International Edition 2002, 41 (17),
3096-3129.
76. Eckert, H., Structural characterization of noncrystalline solids and glasses using
solid state NMR. Progress in Nuclear Magnetic Resonance Spectroscopy 1992, 24
(3), 159-293.
77. Bakhmutov, V. I., Strategies for Solid-State NMR Studies of Materials: From
Diamagnetic to Paramagnetic Porous Solids. Chemical Reviews 2011, 111 (2), 530-
562.
78. Hanna, J. V.; Smith, M. E., Recent technique developments and applications of
solid state NMR in characterising inorganic materials. Solid State Nuclear
Magnetic Resonance 2010, 38 (1), 1-18.
79. Grey, C. P.; Tycko, R., Solid-state NMR in biological and materials physics.
Physics today 2009, 62 (9), 44.
80. Chierotti, M. R.; Gobetto, R., Solid-state NMR studies of weak interactions in
supramolecular systems. Chemical Communications 2008, (14), 1621-1634.
81. Gullion, T.; Schaefer, J., Rotational-echo double-resonance NMR. Journal of
Magnetic Resonance (1969) 1989, 81 (1), 196-200.
82. Fyfe, C. A.; Mueller, K. T.; Grondey, H.; Wong-Moon, K. C., Dipolar dephasing
between quadrupolar and spin-12 nuclei. REDOR and TEDOR NMR experiments
on VPI-5. Chemical Physics Letters 1992, 199 (1–2), 198-204.
83. Hing, A. W.; Vega, S.; Schaefer, J., Transferred-echo double-resonance NMR.
Journal of Magnetic Resonance (1969) 1992, 96 (1), 205-209.
84. van Eck, E. R. H.; Janssen, R.; Maas, W. E. J. R.; Veeman, W. S., A novel
application of nuclear spin-echo double-resonance to aluminophosphates and
aluminosilicates. Chemical Physics Letters 1990, 174 (5), 428-432.
85. Gullion, T., Measurement of dipolar interactions between spin-12 and quadrupolar
nuclei by rotational-echo, adiabatic-passage, double-resonance NMR. Chemical
Physics Letters 1995, 246 (3), 325-330.
86. Duer, M. J., Solid state NMR spectroscopy: principles and applications. John
Wiley & Sons: 2008.
87. Kaupp, M.; Bühl, M.; Malkin, V. G., Calculation of NMR and EPR parameters:
theory and applications. John Wiley & Sons: 2006.

209
88. Wasylishen, R. E.; Ashbrook, S. E.; Wimperis, S., NMR of Quadrupolar Nuclei in
Solid Materials. John Wiley & Sons: 2012.
89. Wasylishen, R. E.; Ashbrook, S. E.; Wimperis, S., NMR of Quadrupolar Nuclei in
Solid Materials. John Wiley & Sons Ltd: Chichester, UK, 2012.
90. Kentgens, A., A practical guide to solid-state NMR of half-integer quadrupolar
nuclei with some applications to disordered systems. Geoderma 1997, 80 (3-4),
271-306.
91. Lapina, O. B.; Khabibulin, D. F.; Shubin, A. A.; Terskikh, V. V., Practical aspects
of 51V and 93Nb solid-state NMR spectroscopy and applications to oxide
materials. Progress in Nuclear Magnetic Resonance Spectroscopy 2008, 53 (3),
128-191.
92. Liu, T., Investigation of Local Structures in Layered Niobates by Solid-state NMR
Spectroscopy. 2012.
93. Herzfeld, J.; Berger, A. E., Sideband intensities in NMR spectra of samples
spinning at the magic angle. The Journal of Chemical Physics 1980, 73 (12), 6021-
6030.
94. Lapina, O. B.; Khabibulin, D. F.; Romanenko, K. V.; Gan, Z.; Zuev, M. G.;
Krasil’nikov, V. N.; Fedorov, V. E., 93Nb NMR chemical shift scale for niobia
systems. Solid State Nuclear Magnetic Resonance 2005, 28 (2–4), 204-224.
95. Hahn, E. L., Spin Echoes. Physical Review 1950, 80 (4), 580-594.
96. Massiot, D.; Farnan, I.; Gautier, N.; Trumeau, D.; Trokiner, A.; Coutures, J. P.,
71Ga and 69Ga nuclear magnetic resonance study of β-Ga2O3: resolution of four-
and six-fold coordinated Ga sites in static conditions. Solid State Nuclear Magnetic
Resonance 1995, 4 (4), 241-248.
97. Bureau, B.; Silly, G.; Buzaré, J. Y.; Legein, C.; Massiot, D., From crystalline to
glassy gallium fluoride materials: an NMR study of 69Ga and 71Ga quadrupolar
nuclei. Solid State Nuclear Magnetic Resonance 1999, 14 (3–4), 181-190.
98. Wang, X., Determination of local structures of layered niobate, niobate nanosheet
and mesoporous niobium oxide by solid state NMR and PXRD. CLARK
UNIVERSITY: 2009.
99. Smith, M. E.; van Eck, E. R. H., Recent advances in experimental solid state NMR
methodology for half-integer spin quadrupolar nuclei. Progress in Nuclear
Magnetic Resonance Spectroscopy 1999, 34 (2), 159-201.
100. Du, L.-S.; Schurko, R. W.; Lim, K. H.; Grey, C. P., A Solid-State 93Nb and 19F
NMR Spectroscopy and X-ray Diffraction Study of Potassium
Heptafluoroniobate(V):  Characterization of 93Nb, 19F Coupling, and Fluorine
Motion. The Journal of Physical Chemistry A 2001, 105 (4), 760-768.
101. Schurko, R. W.; Hung, I.; Macdonald, C. L. B.; Cowley, A. H., Anisotropic NMR
Interaction Tensors in the Decamethylaluminocenium Cation. Journal of the
American Chemical Society 2002, 124 (44), 13204-13214.
102. Lowe, I. J., Free Induction Decays of Rotating Solids. Physical Review Letters
1959, 2 (7), 285-287.

210
103. Andrew, E. R.; Bradbury, A.; Eades, R. G., Removal of Dipolar Broadening of
Nuclear Magnetic Resonance Spectra of Solids by Specimen Rotation. Nature
1959, 183 (4678), 1802-1803.
104. Andrew, E. R.; Eades, R. G., Possibilities for high-resolution nuclear magnetic
resonance spectra of crystals. Discussions of the Faraday Society 1962, 34 (0), 38-
42.
105. Andrew, E. R., The narrowing of NMR spectra of solids by high-speed specimen
rotation and the resolution of chemical shift and spin multiplet structures for solids.
Progress in Nuclear Magnetic Resonance Spectroscopy 1971, 8 (1), 1-39.
106. Laage, S.; Sachleben, J. R.; Steuernagel, S.; Pierattelli, R.; Pintacuda, G.; Emsley,
L., Fast acquisition of multi-dimensional spectra in solid-state NMR enabled by
ultra-fast MAS. Journal of Magnetic Resonance 2009, 196 (2), 133-141.
107. Bertini, I.; Emsley, L.; Lelli, M.; Luchinat, C.; Mao, J.; Pintacuda, G., Ultrafast
MAS Solid-State NMR Permits Extensive 13C and 1H Detection in Paramagnetic
Metalloproteins. Journal of the American Chemical Society 2010, 132 (16), 5558-
5559.
108. Agarwal, V.; Penzel, S.; Szekely, K.; Cadalbert, R.; Testori, E.; Oss, A.; Past, J.;
Samoson, A.; Ernst, M.; Böckmann, A.; Meier, B. H., De Novo 3D Structure
Determination from Sub-milligram Protein Samples by Solid-State 100 kHz MAS
NMR Spectroscopy. Angewandte Chemie International Edition 2014, 53 (45),
12253-12256.
109. Miah, H. K.; Bennett, D. A.; Iuga, D.; Titman, J. J., Measuring proton shift tensors
with ultrafast MAS NMR. Journal of Magnetic Resonance 2013, 235, 1-5.
110. Bak, M.; Rasmussen, J. T.; Nielsen, N. C., SIMPSON: a general simulation
program for solid-state NMR spectroscopy. Journal of Magnetic Resonance 2011,
213 (2), 366-400.
111. O’Dell, L. A.; Rossini, A. J.; Schurko, R. W., Acquisition of ultra-wideline NMR
spectra from quadrupolar nuclei by frequency stepped WURST–QCPMG.
Chemical Physics Letters 2009, 468 (4–6), 330-335.
112. Hung, I.; Gan, Z., On the practical aspects of recording wideline QCPMG NMR
spectra. Journal of Magnetic Resonance 2010, 204 (2), 256-265.
113. O’Dell, L. A.; Schurko, R. W., QCPMG using adiabatic pulses for faster
acquisition of ultra-wideline NMR spectra. Chemical Physics Letters 2008, 464 (1),
97-102.
114. MacGregor, A. W.; O’Dell, L. A.; Schurko, R. W., New methods for the
acquisition of ultra-wideline solid-state NMR spectra of spin-1/2 nuclides. Journal
of Magnetic Resonance 2011, 208 (1), 103-113.
115. Schurko, R. W., Ultra-wideline solid-state NMR spectroscopy. Accounts of
chemical research 2013, 46 (9), 1985-1995.
116. O'Dell, L. A., The WURST kind of pulses in solid-state NMR. Solid State Nuclear
Magnetic Resonance 2013, 55, 28-41.
117. Kupce, E.; Freeman, R., Adiabatic Pulses for Wideband Inversion and Broadband
Decoupling. Journal of Magnetic Resonance, Series A 1995, 115 (2), 273-276.

211
118. Garwood, M.; DelaBarre, L., The return of the frequency sweep: designing
adiabatic pulses for contemporary NMR. Journal of magnetic resonance 2001, 153
(2), 155-177.
119. Bhattacharyya, R.; Frydman, L., Quadrupolar nuclear magnetic resonance
spectroscopy in solids using frequency-swept echoing pulses. The Journal of
chemical physics 2007, 127 (19), 194503.
120. Larsen, F. H.; Jakobsen, H. J.; Ellis, P. D.; Nielsen, N. C., Sensitivity-Enhanced
Quadrupolar-Echo NMR of Half-Integer Quadrupolar Nuclei. Magnitudes and
Relative Orientation of Chemical Shielding and Quadrupolar Coupling Tensors.
The Journal of Physical Chemistry A 1997, 101 (46), 8597-8606.
121. Meiboom, S.; Gill, D., Modified Spin‐Echo Method for Measuring Nuclear
Relaxation Times. Review of Scientific Instruments 1958, 29 (8), 688-691.
122. Harris, K. J.; Lupulescu, A.; Lucier, B. E. G.; Frydman, L.; Schurko, R. W.,
Broadband adiabatic inversion pulses for cross polarization in wideline solid-state
NMR spectroscopy. Journal of Magnetic Resonance 2012, 224, 38-47.
123. Hartmann, S. R.; Hahn, E. L., Nuclear Double Resonance in the Rotating Frame.
Physical Review 1962, 128 (5), 2042-2053.
124. Pines, A.; Gibby, M. G.; Waugh, J. S., Proton‐Enhanced Nuclear Induction
Spectroscopy. A Method for High Resolution NMR of Dilute Spins in Solids. The
Journal of chemical physics 1972, 56 (4), 1776-1777.
125. Pines, A.; Gibby, M. G.; Waugh, J. S., Proton‐enhanced NMR of dilute spins in
solids. The Journal of chemical physics 1973, 59 (2), 569-590.
126. Mueller, K. T.; Sun, B. Q.; Chingas, G. C.; Zwanziger, J. W.; Terao, T.; Pines, A.,
Dynamic-angle spinning of quadrupolar nuclei. Journal of Magnetic Resonance
(1969) 1990, 86 (3), 470-487.
127. Chmelka, B. F.; Mueller, K. T.; Pines, A.; Stebbins, J.; Wu, Y.; Zwanziger, J. W.,
Oxygen-17 NMR in solids by dynamic-angle spinning and double rotation. Nature
1989, 339 (6219), 42-43.
128. Frydman, L.; Harwood, J. S., Isotropic Spectra of Half-Integer Quadrupolar Spins
from Bidimensional Magic-Angle Spinning NMR. Journal of the American
Chemical Society 1995, 117 (19), 5367-5368.
129. Amoureux, J.-P.; Fernandez, C.; Steuernagel, S., ZFiltering in MQMAS NMR.
Journal of Magnetic Resonance, Series A 1996, 123 (1), 116-118.
130. Brown, S. P.; Wimperis, S., Two-dimensional multiple-quantum MAS NMR of
quadrupolar nuclei: a comparison of methods. Journal of Magnetic Resonance
1997, 128 (1), 42-61.
131. Gan, Z.; Kwak, H.-T., Enhancing MQMAS sensitivity using signals from multiple
coherence transfer pathways. Journal of Magnetic Resonance 2004, 168 (2), 346-
351.
132. Massiot, D.; Touzo, B.; Trumeau, D.; Coutures, J. P.; Virlet, J.; Florian, P.;
Grandinetti, P. J., Two-dimensional magic-angle spinning isotropic reconstruction
sequences for quadrupolar nuclei. Solid State Nuclear Magnetic Resonance 1996, 6
(1), 73-83.

212
133. Baltisberger, J. H.; Xu, Z.; Stebbins, J. F.; Wang, S. H.; Pines, A., Triple-Quantum
Two-Dimensional 27Al Magic-Angle Spinning Nuclear Magnetic Resonance
Spectroscopic Study of Aluminosilicate and Aluminate Crystals and Glasses.
Journal of the American Chemical Society 1996, 118 (30), 7209-7214.
134. Prasad, S.; Zhao, P.; Huang, J.; Fitzgerald, J. J.; Shore, J. S., Niobium-93 MQMAS
NMR Spectroscopic Study of Alkali and Lead Niobates. Solid State Nuclear
Magnetic Resonance 2001, 19 (1–2), 45-62.
135. Madhu, P. K.; Goldbourt, A.; Frydman, L.; Vega, S., Sensitivity enhancement of
the MQMAS NMR experiment by fast amplitude modulation of the pulses.
Chemical Physics Letters 1999, 307 (1–2), 41-47.
136. Goldbourt, A.; Madhu, P. K.; Vega, S., Enhanced conversion of triple to single-
quantum coherence in the triple-quantum MAS NMR spectrosocopy of spin-5/2
nuclei. Chemical Physics Letters 2000, 320 (5–6), 448-456.
137. Colaux, H.; Dawson, D. M.; Ashbrook, S. E., Efficient Amplitude-Modulated
Pulses for Triple- to Single-Quantum Coherence Conversion in MQMAS NMR.
The Journal of Physical Chemistry A 2014, 118 (31), 6018-6025.
138. Fernandez, C.; Lang, D. P.; Amoureux, J. P.; Pruski, M., Measurement of
Heteronuclear Dipolar Interactions between Quadrupolar and Spin-1/2 Nuclei in
Solids by Multiple-Quantum REDOR NMR. Journal of the American Chemical
Society 1998, 120 (11), 2672-2673.
139. Oas, T. G.; Griffin, R. G.; Levitt, M. H., Rotary resonance recoupling of dipolar
interactions in solid‐state nuclear magnetic resonance spectroscopy. The Journal of
Chemical Physics 1988, 89 (2), 692-695.
140. Levitt, M. H.; Oas, T. G.; Griffin, R. G., Rotary Resonance Recoupling in
Heteronuclear Spin Pair Systems. Israel Journal of Chemistry 1988, 28 (4), 271-
282.
141. Yoon, K. H.; Cho, Y. S.; Kang, D. H., Molten salt synthesis of lead-based relaxors.
Journal of materials science 1998, 33 (12), 2977-2984.
142. Gopalan, S.; Mehta, K.; Virkar, A. V., Synthesis of oxide perovskite solid solutions
using the molten salt method. Journal of materials research 1996, 11 (08), 1863-
1865.
143. Lee, Y.-h.; Cho, J.-h.; Kim, B.-i.; Choi, D.-k., Piezoelectric properties and
densification based on control of volatile mass of potassium and sodium in (K0.
5Na0. 5) NbO3 ceramics. Japanese Journal of Applied Physics 2008, 47 (6R),
4620.
144. Yang, G.; Kong, Y.; Hou, W.; Yan, Q., Heating Behavior and Crystal Growth
Mechanism in Microwave Field. The Journal of Physical Chemistry B 2005, 109
(4), 1371-1379.
145. Le Bail, A.; Duroy, H.; Fourquet, J., Ab-initio structure determination of LiSbWO
6 by X-ray powder diffraction. Materials Research Bulletin 1988, 23 (3), 447-452.
146. TOPAS, Bruker: Karlsruhe, 2003.
147. Grandinetti, P. RMN, Ohio State University: Columbus, OH, 2005.
148. Eichele, K.; Wasylishen, R. E. WSOLIDS NMR Simulation Package, University of
Tuebingen: Tuebingen, Germany, 2001.

213
149. Massiot, D.; Fayon, F.; Capron, M.; King, I.; Le Calvé, S.; Alonso, B.; Durand, J.
O.; Bujoli, B.; Gan, Z.; Hoatson, G., Modelling one‐and two‐dimensional solid‐
state NMR spectra. Magnetic Resonance in Chemistry 2002, 40 (1), 70-76.
150. Kemp, T. F.; Smith, M. E., QuadFit—a new cross-platform computer program for
simulation of NMR line shapes from solids with distributions of interaction
parameters. Solid state nuclear magnetic resonance 2009, 35 (4), 243-252.
151. Wang, X. Determination of local structures of layered niobate, niobate nanosheet
and mesoporous niobium oxide by solid state NMR and PXRD. 2009.
152. Geselbracht, M. J.; Walton, R. I.; Cowell, E. S.; Millange, F.; O’Hare, D., A
furnace for the in situ study of the formation of inorganic solids at high temperature
using time-resolved energy-dispersive x-ray diffraction. Review of Scientific
Instruments 2000, 71 (11), 4177-4181.
153. Takano, Y.; Kimishima, Y.; Yamadaya, T.; Ogawa, S.; Môri, N., High Pressure
Studies of the Non-Copper Superconductors KCa2Nb3O10 and RbCa2Nb3O10.
The Review of High Pressure Science and Technology 1998, 7, 589-591.
154. Kurzman, J. A.; Geselbracht, M. J. In Probing octahedral tilting in Dion-Jacobson
layered perovskites with neutron powder diffraction and Raman spectroscopy,
Materials Research Society Symposium Proceedings, Warrendale, Pa.; Materials
Research Society; 1999: 2007; p 74.
155. Kulischow, N.; Ladasiu, C.; Marschall, R., Layered Dion-Jacobson type niobium
oxides for photocatalytic hydrogen production prepared via molten salt synthesis.
Catalysis Today 2016, (Copyright (C) 2017 American Chemical Society (ACS).
All Rights Reserved.), Ahead of Print.
156. Sirotinkin, S.; Sirotinkin, V.; Trunov, V., Structure of low-temperature
modification of BaNb 2 O 6. Zhurnal Neorganicheskoj Khimii 1990, 35 (6), 1609-
1611.
157. Fieller, E. C., Some problems in interval estimation. Journal of the Royal Statistical
Society. Series B (Methodological) 1954, 175-185.
158. Ziegler, C. Two-dimensional transition metal oxide nanosheets for
nanoarchitectonics. Ludwig-Maximilians-Universität, 2015.
159. Schaak, R. E.; Mallouk, T. E., Topochemical synthesis of three-dimensional
perovskites from lamellar precursors. Journal of the American Chemical Society
2000, 122 (12), 2798-2803.
160. Oral, A.; Mecartney, M., Properties of sol-gel derived strontium barium niobate
ceramics and the effect of V2O5 additive. Journal of Materials Science 2001, 36
(22), 5519-5527.
161. Jehng, J. M.; Wachs, I. E., Molecular structures of supported niobium oxide
catalysts under in situ conditions. The Journal of Physical Chemistry 1991, 95 (19),
7373-7379.
162. Jehng, J. M.; Wachs, I. E., Structural chemistry and Raman spectra of niobium
oxides. Chemistry of Materials 1991, 3 (1), 100-107.
163. Rietveld, H., A profile refinement method for nuclear and magnetic structures.
Journal of Applied Crystallography 1969, 2 (2), 65-71.

214
164. Kato, H.; Kudo, A., Energy structure and photocatalytic activity for water splitting
of Sr2(Ta1−XNbX)2O7 solid solution. Journal of Photochemistry and
Photobiology A: Chemistry 2001, 145 (1–2), 129-133.
165. Hojamberdiev, M.; Bekheet, M. F.; Zahedi, E.; Wagata, H.; Vequizo, J. J. M.;
Yamakata, A.; Yubuta, K.; Gurlo, A.; Domen, K.; Teshima, K., The contrasting
effect of the Ta/Nb ratio in (111)-layered B-site deficient hexagonal perovskite Ba
5 Nb 4− x Ta x O 15 crystals on visible-light-induced photocatalytic water
oxidation activity of their oxynitride derivatives. Dalton Transactions 2016, 45
(31), 12559-12568.
166. Weitzel, H., Kristallstrukturverfeinerung von Wolframiten und Columbiten.
Zeitschrift für Kristallographie-Crystalline Materials 1976, 144 (1-6), 238-258.
167. Sugimoto, W.; Ohkawa, H.; Naito, M.; Sugahara, Y.; Kuroda, K., Synthesis and
Structures of Reduced Niobates with Four Perovskite-like Layers and Their
Semiconducting Properties. Journal of Solid State Chemistry 1999, 148 (2), 508-
512.
168. Liang, Z.-H.; Tang, K.-B.; Chen, Q.-W.; Zheng, H.-G., RbCa2Nb3O10 from X-ray
powder data. Acta Crystallographica Section E: Structure Reports Online 2009, 65
(6), i44-i44.
169. Wang, X.; Adhikari, J.; Smith, L. J., An Investigation of Distortions of the
Dion−Jacobson Phase RbSr2Nb3O10 and Its Acid-Exchanged Form with 93Nb
Solid State NMR and DFT Calculations. The Journal of Physical Chemistry C
2009, 113 (40), 17548-17559.
170. Ratheesh, R.; Wöhlecke, M.; Berge, B.; Wahlbrink, T.; Haeuseler, H.; Rühl, E.;
Blachnik, R.; Balan, P.; Santha, N.; Sebastian, M., Raman study of the ordering in
Sr (B 0.5′ Nb 0.5) O 3 compounds. Journal of Applied Physics 2000, 88 (5), 2813-
2818.
171. Lufaso, M. W., Crystal structures, modeling, and dielectric property relationships
of 2: 1 ordered Ba3MM'2O9 (M= Mg, Ni, Zn; M'= Nb, Ta) perovskites. Chemistry
of Materials 2004, 16 (11), 2148-2156.
172. Welk, M. E.; Norquist, A. J.; Arnold, F. P.; Stern, C. L.; Poeppelmeier, K. R., Out-
of-center distortions in d0 transition metal oxide fluoride anions. Inorganic
chemistry 2002, 41 (20), 5119-5125.
173. Koller, H.; Engelhardt, G.; Kentgens, A. P. M.; Sauer, J., 23Na NMR Spectroscopy
of Solids: Interpretation of Quadrupole Interaction Parameters and Chemical Shifts.
The Journal of Physical Chemistry 1994, 98 (6), 1544-1551.
174. Brown, I. t.; Shannon, R., Empirical bond-strength-bond-length curves for oxides.
Acta Crystallographica Section A: Crystal Physics, Diffraction, Theoretical and
General Crystallography 1973, 29 (3), 266-282.
175. Le Bail, A., Whole powder pattern decomposition methods and applications: a
retrospection. Powder Diffraction 2005, 20 (04), 316-326.
176. Duer, M. J., Solid-state NMR spectroscopy: Principles and Applications. Blackwell
Science: Oxford, 2002.
177. Kappe, C. O., Microwave dielectric heating in synthetic organic chemistry.
Chemical Society Reviews 2008, 37 (6), 1127-1139.

215
178. Gabriel, C.; Gabriel, S.; H. Grant, E.; S. J. Halstead, B.; Michael P. Mingos, D.,
Dielectric parameters relevant to microwave dielectric heating. Chemical Society
Reviews 1998, 27 (3), 213-224.
179. Tauc, J.; Grigorovici, R.; Vancu, A., Optical properties and electronic structure of
amorphous germanium. physica status solidi (b) 1966, 15 (2), 627-637.
180. Tauc, J., Optical Properties and Electronic Structure of Amorphous
Semiconductors. In Optical Properties of Solids: Papers from the NATO Advanced
Study Institute on Optical Properties of Solids Held August 7–20, 1966, at
Freiburg, Germany, Nudelman, S.; Mitra, S. S., Eds. Springer US: Boston, MA,
1969; pp 123-136.
181. Davis, E.; Mott, N., Conduction in non-crystalline systems V. Conductivity, optical
absorption and photoconductivity in amorphous semiconductors. Philosophical
Magazine 1970, 22 (179), 0903-0922.
182. Maeda, K.; Mallouk, T. E., Comparison of two- and three-layer restacked Dion-
Jacobson phase niobate nanosheets as catalysts for photochemical hydrogen
evolution. J Mater Chem 2009, 19 (27), 4813-4818.
183. Akbarian-Tefaghi, S.; Teixeira Veiga, E.; Amand, G.; Wiley, J. B., Rapid
Topochemical Modification of Layered Perovskites via Microwave Reactions.
Inorganic Chemistry 2016, 55 (4), 1604-1612.
184. Wang, Y.; Delahaye, E.; Leuvrey, C. d.; Leroux, F.; Rabu, P.; Rogez, G., Efficient
Microwave-Assisted Functionalization of the Aurivillius-Phase Bi2SrTa2O9.
Inorganic chemistry 2016, 55 (8), 4039-4046.
185. Wang, Y.; Delahaye, E.; Leuvrey, C.; Leroux, F.; Rabu, P.; Rogez, G., Post-
Synthesis Modification of the Aurivillius Phase Bi2SrTa2O9 via In Situ
Microwave-Assisted “Click Reaction”. Inorganic Chemistry 2016, 55 (19), 9790-
9797.
186. Osada, M.; Sasaki, T., Two‐Dimensional Dielectric Nanosheets: Novel
Nanoelectronics From Nanocrystal Building Blocks. Advanced Materials 2012, 24
(2), 210-228.
187. Ziegler, C.; Werner, S.; Bugnet, M.; Wörsching, M.; Duppel, V.; Botton, G. A.;
Scheu, C.; Lotsch, B. V., Artificial solids by design: assembly and electron
microscopy study of nanosheet-derived heterostructures. Chemistry of Materials
2013, 25 (24), 4892-4900.
188. Osada, M.; Sasaki, T., Chemical nanomanipulation of two-dimensional nanosheets
and its applications. 2011.
189. Xu, F.; Ebina, Y.; Bando, Y.; Sasaki, T., Structural characterization of (TBA, H)
Ca2Nb3O10 nanosheets formed by delamination of a precursor-layered perovskite.
The Journal of Physical Chemistry B 2003, 107 (36), 9638-9645.
190. Zhou, K. G.; Mao, N. N.; Wang, H. X.; Peng, Y.; Zhang, H. L., A Mixed‐Solvent
Strategy for Efficient Exfoliation of Inorganic Graphene Analogues. Angewandte
Chemie 2011, 123 (46), 11031-11034.
191. Štengl, V.; Henych, J., Strongly luminescent monolayered MoS 2 prepared by
effective ultrasound exfoliation. Nanoscale 2013, 5 (8), 3387-3394.

216
192. Coleman, J. N.; Lotya, M.; O’Neill, A.; Bergin, S. D.; King, P. J.; Khan, U.;
Young, K.; Gaucher, A.; De, S.; Smith, R. J., Two-dimensional nanosheets
produced by liquid exfoliation of layered materials. Science 2011, 331 (6017), 568-
571.
193. Han, W.-Q.; Wu, L.; Zhu, Y.; Watanabe, K.; Taniguchi, T., Structure of chemically
derived mono- and few-atomic-layer boron nitride sheets. Applied Physics Letters
2008, 93 (22), 223103.
194. Lin, Y.; Williams, T. V.; Connell, J. W., Soluble, Exfoliated Hexagonal Boron
Nitride Nanosheets. The Journal of Physical Chemistry Letters 2010, 1 (1), 277-
283.
195. Zhi, C.; Bando, Y.; Tang, C.; Kuwahara, H.; Golberg, D., Large-Scale Fabrication
of Boron Nitride Nanosheets and Their Utilization in Polymeric Composites with
Improved Thermal and Mechanical Properties. Advanced Materials 2009, 21 (28),
2889-2893.
196. Ebina, Y.; Sakai, N.; Sasaki, T., Photocatalyst of lamellar aggregates of RuO x-
loaded perovskite nanosheets for overall water splitting. The Journal of Physical
Chemistry B 2005, 109 (36), 17212-17216.
197. Hansen, C. M., The three dimensional solubility parameter. Danish Technical:
Copenhagen 1967, 14.
198. Agrawal, K. V.; Topuz, B.; Jiang, Z.; Nguenkam, K.; Elyassi, B.; Francis, L. F.;
Tsapatsis, M.; Navarro, M., Solution‐processable exfoliated zeolite nanosheets
purified by density gradient centrifugation. AIChE Journal 2013, 59 (9), 3458-
3467.
199. Keller, A. A.; Wang, H.; Zhou, D.; Lenihan, H. S.; Cherr, G.; Cardinale, B. J.;
Miller, R.; Ji, Z., Stability and aggregation of metal oxide nanoparticles in natural
aqueous matrices. Environmental science & technology 2010, 44 (6), 1962-1967.
200. Marinuc, M.; Rus, F., THE EFFECT OF SOLID PARTICLE SIZE UPON TIME
AND SEDIMENTATION RATE. Bulletin of the Transilvania University of
Brasov, Series II. Forestry, Wood Industry, Agricultural Food Engineering 2012,
(1).
201. Ida, S.; Ogata, C.; Eguchi, M.; Youngblood, W. J.; Mallouk, T. E.; Matsumoto, Y.,
Photoluminescence of Perovskite Nanosheets Prepared by Exfoliation of Layered
Oxides, K2Ln2Ti3O10, KLnNb2O7, and RbLnTa2O7 (Ln: Lanthanide Ion).
Journal of the American Chemical Society 2008, 130 (22), 7052-7059.
202. Teagarden, D. L.; Baker, D. S., Practical aspects of lyophilization using non-
aqueous co-solvent systems. European Journal of Pharmaceutical Sciences 2002,
15 (2), 115-133.
203. Yong, Z.; Yingjie, D.; Xueli, W.; Jinghua, X.; Zhengqiang, L., Conformational and
bioactivity analysis of insulin: Freeze-drying TBA/water co-solvent system in the
presence of surfactant and sugar. International journal of pharmaceutics 2009, 371
(1), 71-81.
204. Ebina, Y.; Sasaki, T.; Watanabe, M., Study on exfoliation of layered perovskite-
type niobates. Solid State Ionics 2002, 151 (1), 177-182.

217
205. Friedrich, W.; Knipping, P.; Laue, M., Interferenzerscheinungen bei
roentgenstrahlen. Annalen der Physik 1913, 346 (10), 971-988.

218
Appendix A: Indexed Lists of XRD Pattern hkl Reflections
Figure A.1. List of indexed hkl reflections obtained from XRD pattern of RbSr2Nb3O10 microwave
sample.
h k l d (Å) 2θ Intensity
0 0 2 15.16297 5.82390 36.9
0 2 0 3.88017 22.90106 11.5
2 0 4 3.46316 25.70316 17
1 0 9 3.09239 28.84789 29.6
0 1 9 3.09077 28.86336 38.8
2 0 6 3.08422 28.92597 141
0 2 6 3.07780 28.98763 27.5
0 0 10 3.03259 29.42942 10.2
1 2 5 3.01376 29.61757 20.6
0 1 10 2.82458 31.65144 30.7
2 2 0 2.74824 32.55477 116
2 2 1 2.73702 32.69191 13.7
2 0 8 2.71591 32.95325 36.6
0 3 1 2.57742 34.77877 10.7
0 2 9 2.54414 35.24846 16.4
3 1 2 2.42957 36.96945 12.5
0 1 12 2.40296 37.39391 12.2
3 1 4 2.34108 38.42045 11.9
2 2 7 2.32068 38.77158 13.1
0 3 6 2.30272 39.08630 17.3
0 2 11 2.24739 40.08929 11.4
0 0 14 2.16614 41.66153 11.3
2 2 9 2.12970 42.40841 11.3
1 3 8 2.06052 43.90495 11.4
0 3 9 2.05187 44.09970 12.5
1 2 12 2.04340 44.29221 16.2
2 3 5 2.03023 44.59473 15.7
1 1 14 2.01529 44.94347 21.3
0 2 13 1.99927 45.32348 18
4 0 0 1.94653 46.62291 47.5
2 2 11 1.94635 46.62732 27.6
0 4 0 1.94008 46.78696 17.3
1 1 15 1.89744 47.90337 12.2
1 4 2 1.86818 48.70194 18.2
2 4 3 1.71135 53.50172 19.1
4 1 9 1.64710 55.76625 41.2
4 2 6 1.64513 55.83892 18.3
3 2 13 1.58383 58.20197 13.5

219
Figure A.2. List of indexed hkl reflections obtained from XRD pattern of RbSr2Nb3O10 molten salt
sample
h k l d (Å) 2θ Intensity
0 0 2 15.26925 5.78333 87.5
0 0 4 7.63462 11.58146 7.35
2 0 0 3.89776 22.79630 8.71
0 2 0 3.88346 22.88136 16.4
0 0 8 3.81731 23.28339 8.14
1 0 7 3.80702 23.34718 10.2
2 0 4 3.47151 25.64029 8.32
0 2 4 3.46140 25.71649 7.82
2 1 1 3.46125 25.71761 7.45
2 0 5 3.28570 27.11719 8.21
1 1 8 3.13639 28.43462 15.6
1 0 9 3.11122 28.66960 26.2
0 1 9 3.10939 28.68680 6.2
2 0 6 3.09457 28.82709 51
0 2 6 3.08740 28.89550 239
0 0 10 3.05385 29.22001 69.3
2 1 5 3.02606 29.49437 7.88
1 2 5 3.02103 29.54463 10.3
1 0 10 2.84345 31.43596 6.58
2 2 0 2.75107 32.52039 200
2 2 1 2.73997 32.65576 16.5
2 0 8 2.72725 32.81241 23
0 2 8 2.72233 32.87329 5.28
2 1 7 2.72226 32.87419 5.07
1 2 7 2.71860 32.91978 12.8
2 2 2 2.70747 33.05890 9.16
0 0 12 2.54487 35.23801 9.89
0 0 14 2.18132 41.35822 29.2
2 1 11 2.17113 41.56136 6.3
0 2 12 2.12855 42.43237 10.4
2 2 10 2.04398 44.27875 10.4
2 2 11 1.95413 46.43085 12
2 1 13 1.94768 46.59378 19.6
1 2 13 1.94633 46.62786 48.2
4 0 1 1.94493 46.66362 40.9
2 0 14 1.90351 47.74113 9.93
2 2 12 1.86814 48.70299 18
4 1 7 1.73447 52.73308 7.08
2 2 14 1.70923 53.57335 17.8
1 1 17 1.70767 53.62610 10.7
3 1 13 1.70032 53.87658 6.86
1 0 18 1.65778 55.37609 10.2
4 1 9 1.65133 55.61093 7.79
4 2 6 1.64801 55.73264 11.4
1 4 9 1.64725 55.76084 20
3 2 12 1.64663 55.78342 40.2
2 3 12 1.64528 55.83332 12.2
2 4 6 1.64476 55.85242 6.12

220
4 0 10 1.64285 55.92315 8.79
3 0 15 1.60260 57.45652 6.07
2 3 13 1.58865 58.00872 6.48
4 2 8 1.58467 58.16831 14.7
2 4 8 1.58178 58.28482 8.68

Figure A.3. List of indexed hkl reflections obtained from XRD pattern of RbCa0.3Sr1.7Nb3O10
h k l d (Å) 2θ Intensity
0 0 2 14.78497 5.97293 17.5
0 2 1 3.84304 23.12534 15.7
2 0 2 3.73513 23.80306 10.5
0 1 7 3.70928 23.97145 6.17
2 0 4 3.42188 26.01861 14.4
1 1 8 3.06265 29.13422 44.8
0 2 6 3.04660 29.29107 34.2
2 0 6 3.03903 29.36572 31.2
2 2 0 2.73517 32.71471 15.3
2 2 1 2.72354 32.85831 49
2 2 2 2.68953 33.28582 7.63
2 2 7 2.29592 39.20686 5.21
2 3 0 2.14730 42.04420 7.81
0 3 9 2.03107 44.57539 6.02
0 4 0 1.93796 46.84138 8.14
0 4 1 1.93381 46.94790 9.91
4 0 0 1.93018 47.04155 24.1
3 1 10 1.88314 48.29036 6.38
2 2 12 1.83076 49.76427 6.77
3 3 4 1.77038 51.58363 5.82
4 2 4 1.68245 54.49616 12.7
0 0 18 1.64277 55.92593 12.4
0 3 14 1.63532 56.20325 9.83
3 3 8 1.63528 56.20482 9.45
4 1 10 1.58228 58.26454 9.96
1 1 18 1.57336 58.62701 5.87

Figure A.4. List of indexed hkl reflections obtained from XRD pattern of RbCa0.4Sr1.6Nb3O10
h k l d (Å) 2θ Intensity
0 0 2 14.92165 5.91817 96.8
1 0 1 7.59071 11.64870 10.6
0 1 1 7.49886 11.79187 9.22
0 1 3 6.11239 14.47960 6.59
1 1 1 5.42200 16.33517 6.1
0 2 0 3.87372 22.93972 24.1
2 0 2 3.79535 23.42000 30
0 2 2 3.74943 23.71097 6.75
2 0 4 3.47324 25.62732 5.25
1 2 1 3.45039 25.79992 30.9
2 1 3 3.30236 26.97778 9.29
2 1 4 3.16932 28.13302 6.61

221
1 2 4 3.14909 28.31750 5.36
1 1 8 3.08971 28.87350 10.3
2 0 6 3.08091 28.95772 351
0 1 9 3.04844 29.27301 16.4
2 2 1 2.74519 32.59188 190
2 2 2 2.71100 33.01464 45.8
1 2 8 2.54219 35.27650 10.6
3 0 3 2.53024 35.44863 6
2 2 6 2.41126 37.26035 5.74
1 3 4 2.33037 38.60401 6.21
3 2 1 2.16241 41.73679 12.9
1 3 7 2.12625 42.48063 5.31
3 2 5 2.03783 44.41971 11
0 3 9 2.03746 44.42801 10.9
4 0 2 1.94546 46.65000 102
0 4 0 1.93686 46.86955 29.2
4 1 0 1.90215 47.77746 5.4
4 1 1 1.89830 47.88047 5.24
4 0 4 1.89768 47.89714 5.29
4 1 2 1.88688 48.18850 5.25
1 4 2 1.86569 48.77114 10.6
0 0 16 1.86521 48.78469 16.8
4 2 2 1.73853 52.60052 9.11
2 4 3 1.71096 53.51479 23.2
4 2 4 1.70417 53.74506 30.4
3 2 11 1.69371 54.10400 6.11
2 1 16 1.64615 55.80127 61.3
0 3 14 1.64396 55.88210 49.1
1 2 16 1.64329 55.90668 18.3
3 0 15 1.58366 58.20884 16.6

Figure A.5. List of indexed hkl reflections obtained from XRD pattern of RbCa0.5Sr1.5Nb3O10.
h k l d (Å) 2θ Intensity
0 0 2 14.91646 5.92023 48.9
0 2 1 3.85978 23.02366 69.4
0 2 2 3.76637 23.60279 26.9
2 0 2 3.74365 23.74810 26.3
0 1 7 3.73833 23.78244 9.8
1 0 7 3.73273 23.81859 12.7
0 0 8 3.72912 23.84204 8.12
0 2 4 3.45079 25.79686 36
2 1 2 3.37383 26.39584 11.4
2 1 3 3.27087 27.24250 81.6
1 2 4 3.15140 28.29639 15.4
2 1 4 3.14137 28.38857 9.69
1 1 8 3.08424 28.92577 15.1
0 2 6 3.06499 29.11146 109
2 0 6 3.05271 29.23118 47.6

222
0 1 9 3.04982 29.25951 74.6
1 0 9 3.04678 29.28934 90
2 1 5 2.99547 29.80260 27.8
0 0 10 2.98329 29.92704 27.8
0 2 7 2.87414 31.09181 17.4
0 1 10 2.78575 32.10446 8.34
2 2 0 2.74351 32.61246 15.3
2 2 1 2.73198 32.75394 210
0 2 8 2.69279 33.24436 25.2
1 3 4 2.33640 38.50045 37.2
1 1 12 2.26449 39.77380 9.1
2 3 1 2.14926 42.00403 11
2 3 2 2.13272 42.34539 8.55
1 1 14 1.98639 45.63393 9.56
0 4 0 1.94625 46.62998 116
4 0 0 1.93372 46.95022 16.1
1 4 1 1.88365 48.27636 9.15
2 2 12 1.84223 49.43370 16.5
2 4 4 1.69313 54.12411 11.8
2 3 11 1.68715 54.33173 8.42
1 2 16 1.64320 55.90999 38.8
3 3 8 1.64213 55.94991 57.5
2 1 16 1.64178 55.96282 9.81

Figure A.7. List of indexed hkl reflections obtained from XRD pattern of RbCa0.7Sr1.3Nb3O10
h k l d (Å) 2θ Intensity
0 0 2 14.99434 5.88946 45.5
1 1 1 5.40788 16.37811 11.8
0 1 6 4.19585 21.15739 10.5
1 1 5 4.05286 21.91292 11.5
0 2 0 3.86086 23.01712 33
2 0 2 3.78808 23.46563 16.2
0 2 2 3.73891 23.77868 11.1
1 1 6 3.69833 24.04344 10.2
0 2 3 3.60157 24.69949 15
1 2 1 3.43994 25.87966 16.3
0 2 4 3.43246 25.93708 17.9
2 1 3 3.29655 27.02627 10.2
1 2 3 3.27204 27.23259 10.4
2 1 4 3.16539 28.16872 15.1
1 2 4 3.14367 28.36738 5.15
2 0 6 3.08209 28.94640 75.3
1 0 9 3.06601 29.10154 54.1
0 1 9 3.05939 29.16596 36
0 2 6 3.05543 29.20453 58.5
1 2 5 2.99887 29.76797 8.75
0 0 10 2.99887 29.76799 5.43
0 2 7 2.86803 31.15963 11.8

223
2 1 6 2.86249 31.22149 6.01
0 1 10 2.79545 31.99007 6.56
2 2 0 2.74901 32.54540 11.1
2 2 1 2.73753 32.68567 110
2 0 8 2.70760 33.05736 5.24
2 1 7 2.70668 33.06886 6.01
2 2 2 2.70394 33.10331 8.61
1 2 7 2.69306 33.24088 5.27
2 2 3 2.65061 33.78920 5.68
2 1 8 2.55507 35.09281 5.04
3 1 1 2.46425 36.43061 5.93
1 3 4 2.32467 38.70239 9.78
3 2 1 2.15671 41.85227 6.82
1 1 14 1.99592 45.40382 6.25
4 0 2 1.94107 46.76193 22.9
1 0 15 1.93710 46.86333 17.3
2 2 11 1.93574 46.89815 16.3
0 1 15 1.93543 46.90629 16.1
0 4 0 1.93043 47.03495 5.97
2 2 12 1.84917 49.23582 10.5
4 2 4 1.70044 53.87247 5.02
1 1 17 1.67970 54.59267 7.29
3 3 8 1.64644 55.79055 5.38
4 0 10 1.63921 56.05806 40.3
2 4 6 1.63602 56.17729 11.7
3 2 12 1.63518 56.20873 6.41
1 4 9 1.63360 56.26785 5.13
0 4 10 1.62320 56.66076 7.51

Figure A.8. List of indexed hkl reflections obtained from XRD pattern of RbCa1.1Sr0.9Nb3O10
h k l d (Å) 2θ Intensity
0 0 2 15.09998 5.84822 115
0 1 3 6.15008 14.39038 14.9
1 1 1 5.42854 16.31535 6.33
0 1 4 5.41422 16.35882 6.31
0 2 0 3.88420 22.87697 49.8
2 0 2 3.79449 23.42541 72.6
1 2 1 3.45763 25.74497 27.7
0 2 4 3.45392 25.77310 17.8
1 2 3 3.28945 27.08564 96
1 0 9 3.08491 28.91940 137
0 1 9 3.08046 28.96207 279
0 2 6 3.07504 29.01420 60.7
2 1 5 3.02823 29.47281 39.7
1 2 5 3.01566 29.59846 12.3
0 2 7 2.88665 30.95358 7.35
1 1 9 2.86711 31.16990 5.7
1 2 6 2.86274 31.21868 11.8
0 1 10 2.81478 31.76459 5.91
2 2 1 2.74777 32.56049 312

224
2 0 8 2.71924 32.91179 6.84
2 1 7 2.71799 32.92728 12.7
2 2 2 2.71427 32.97373 26.5
1 2 7 2.70889 33.04106 5.79
2 2 3 2.66106 33.65245 5.26
2 1 8 2.56654 34.93087 5.51
0 2 9 2.53923 35.31892 10.3
3 0 3 2.52966 35.45696 13.9
0 0 12 2.51666 35.64619 6.01
3 1 1 2.46880 36.36112 5.82
1 1 11 2.45805 36.52576 6.77
3 1 3 2.40534 37.35542 9.46
2 2 7 2.32448 38.70577 64.5
1 3 6 2.20931 40.81075 9.96
3 2 1 2.16280 41.72889 35.6
3 2 3 2.11975 42.61725 15.3
1 2 12 2.03938 44.38395 5.63
2 2 10 2.03703 44.43807 5.97
2 3 5 2.03441 44.49818 7.81
1 1 14 2.00910 45.08950 7.11
3 2 6 1.99143 45.51186 5.18
2 3 6 1.98546 45.65639 7.97
0 3 10 1.96578 46.13976 9.72
2 2 11 1.94617 46.63189 23.6
4 0 2 1.94383 46.69138 113
0 4 0 1.94210 46.73552 35.1
4 1 0 1.90057 47.81954 23.8
4 0 4 1.89724 47.90871 17.5
4 1 1 1.89682 47.92006 14.8
4 1 3 1.86758 48.71863 11
2 2 12 1.85940 48.94706 22.2
1 4 5 1.79952 50.68870 6.36
2 2 13 1.77710 51.37455 6.02
0 4 7 1.77094 51.56623 5.71
2 4 3 1.71482 53.38467 10.6
4 2 4 1.70475 53.72543 43.2
2 4 6 1.64473 55.85370 63.7
4 0 10 1.64418 55.87396 55.7
1 4 9 1.64353 55.89801 30.3
3 2 12 1.64272 55.92795 13.4
1 4 10 1.59915 57.59194 5.54
1 2 17 1.58228 58.26450 5.79
2 3 13 1.58212 58.27089 6.46
2 4 8 1.58041 58.34016 26.6

225
Figure A.9. List of indexed hkl reflections obtained from XRD pattern of RbCa1.4Sr0.6Nb3O10
h k l d (Å) 2θ Intensity
0 0 2 14.88440 5.93300 105
0 0 4 7.44220 11.88197 13.3
0 1 3 6.09310 14.52569 13
0 2 0 3.85994 23.02269 90.4
2 0 2 3.76066 23.63916 20.1
0 2 2 3.73635 23.79519 25
1 0 7 3.73087 23.83065 17.1
0 1 7 3.72490 23.86945 21.1
0 0 8 3.72110 23.89415 33.6
2 1 1 3.44822 25.81645 10.4
0 2 4 3.42649 25.98303 81.6
1 1 8 3.07804 28.98529 97.5
2 0 6 3.05968 29.16305 173
0 2 6 3.04655 29.29158 249
1 0 9 3.04358 29.32083 71.6
2 2 0 2.73883 32.66980 96.4
2 2 1 2.72731 32.81166 207
2 2 2 2.69360 33.23403 16.6
1 2 7 2.68260 33.37440 16.6
0 2 8 2.67896 33.42107 17.4
2 1 11 2.13436 42.31137 11.4
1 2 11 2.13100 42.38125 10.2
4 0 0 1.94338 46.70289 36.6
0 4 0 1.92997 47.04684 154
4 0 4 1.88033 48.36710 10.7
1 1 15 1.86590 48.76543 11.9
2 0 14 1.86544 48.77827 10.2
1 4 3 1.84060 49.48031 16.4
2 4 3 1.70295 53.78667 11
4 2 4 1.69042 54.21781 21.4
2 3 11 1.68132 54.53556 10.4
4 1 8 1.68126 54.53789 10.3
2 1 16 1.63989 56.03308 17.8
3 3 8 1.63918 56.05927 10.4
0 3 14 1.63915 56.06052 10.2
4 1 9 1.63745 56.12388 14.3
2 4 6 1.63236 56.31430 86.1
1 4 9 1.62990 56.40688 24.3
3 2 13 1.56794 58.84947 10.1
2 4 8 1.56770 58.85934 10.1

226
Figure A.10. List of indexed hkl reflections obtained from XRD pattern of RbCa1.9Sr0.1Nb3O10
h k l d (Å) 2θ Intensity
0 0 2 14.86902 5.93914 65.3
0 0 4 7.43451 11.89431 11.7
0 0 6 4.95634 17.88193 7.92
2 0 0 3.85350 23.06170 83.1
0 2 2 3.74482 23.74063 10.9
2 0 2 3.73027 23.83458 47.8
0 1 7 3.72409 23.87470 21.6
1 0 7 3.72050 23.89810 11.4
2 1 1 3.42656 25.98250 47.8
2 0 4 3.42123 26.02365 15
1 2 3 3.26516 27.29107 6.19
0 2 5 3.24350 27.47686 5.03
1 2 4 3.13554 28.44252 6.46
1 1 8 3.07291 29.03481 36.6
0 2 6 3.05007 29.25700 308
2 0 6 3.04220 29.33445 62.4
0 1 9 3.03884 29.36755 5.25
0 0 10 2.97380 30.02477 6.42
0 2 7 2.86073 31.24121 7.86
1 2 6 2.83606 31.52006 7.35
0 1 10 2.77592 32.22126 5.62
2 2 0 2.73049 32.77229 215
2 2 2 2.68559 33.33614 20.5
1 2 7 2.68193 33.38287 9.59
0 2 8 2.68072 33.39844 7.27
1 1 10 2.61168 34.30833 6.5
2 2 4 2.56309 34.97942 5.37
3 0 3 2.48684 36.08823 6.37
2 2 7 2.29697 39.18816 6.2
0 2 11 2.21616 40.67893 10.2
0 3 7 2.20501 40.89394 6.69
2 3 0 2.14369 42.11846 6.95
2 3 1 2.13814 42.23298 6.21
3 2 1 2.13474 42.30341 14
1 2 11 2.12986 42.40513 6.5
2 1 11 2.12784 42.44726 7.81
3 0 9 2.02813 44.64344 9.35
2 3 5 2.01669 44.91037 8.58
2 2 10 2.01128 45.03799 8.27
0 4 0 1.93477 46.92306 29.1
0 4 1 1.93069 47.02822 109
1 4 1 1.87282 48.57349 8.78
0 4 4 1.87241 48.58491 16.8
4 1 0 1.86968 48.66042 9.26
0 3 11 1.86634 48.75309 7.65
4 1 1 1.86599 48.76274 6.03
4 0 4 1.86513 48.78672 11.2
4 2 0 1.72477 53.05276 11.1
2 4 2 1.71750 53.29495 5.73

227
1 4 7 1.71654 53.32700 12.9
0 4 8 1.71622 53.33768 17.9
2 4 4 1.68412 54.43734 31.7
4 2 4 1.68015 54.57697 9.25
2 3 11 1.67971 54.59243 10.6
0 0 18 1.65211 55.58231 6.34
0 3 14 1.63979 56.03667 15.3
3 3 8 1.63483 56.22161 20.6
2 4 6 1.63258 56.30618 52.4
1 4 9 1.63176 56.33706 33.6
0 4 11 1.57336 58.62689 7.64
3 0 15 1.56952 58.78448 13.3
2 4 8 1.56777 58.85663 8.42

Figure A.11. List of indexed hkl reflections obtained from XRD pattern of RbCa2Nb3O10
h k l d (Å) 2θ Intensity
0 0 2 15.00294 5.88608 18.7
0 0 6 5.00098 17.72101 10.7
0 2 0 3.89119 22.83530 58.3
2 0 0 3.86377 22.99958 12.7
0 2 2 3.76657 23.60154 7.64
0 1 7 3.75468 23.67738 36.3
0 0 8 3.75074 23.70261 12.2
1 0 7 3.74847 23.71717 10.5
2 0 2 3.74168 23.76081 19.1
0 2 4 3.45413 25.77148 8.57
1 2 1 3.45236 25.78494 37.3
2 1 1 3.43794 25.89502 28.3
1 1 8 3.09580 28.81539 5.7
0 2 6 3.07106 29.05261 65.5
0 1 9 3.06461 29.11518 155
1 0 9 3.06123 29.14803 88.5
2 0 6 3.05753 29.18407 61.5
2 1 5 2.99795 29.77729 11.5
2 1 6 2.84578 31.40958 5.47
2 2 0 2.74174 32.63406 180
2 1 7 2.69270 33.24553 23.4
2 0 8 2.69124 33.26402 15.7
0 3 8 2.13355 42.32828 25.7
2 0 12 2.09924 43.05427 14
3 2 5 2.02226 44.78012 24.1
0 3 10 1.96242 46.22337 5.17
0 1 15 1.93741 46.85532 31.4
1 0 15 1.93656 46.87723 88.3
2 2 11 1.93376 46.94917 8.12
4 1 1 1.87133 48.61473 10.4
4 0 4 1.87084 48.62825 24.8
2 3 8 1.86772 48.71488 24.8

228
3 2 8 1.86390 48.82121 9.07
4 1 3 1.84288 49.41507 25.1
3 3 3 1.79805 50.73305 5
3 0 13 1.71898 53.24528 12.6
4 2 2 1.71897 53.24583 12.4
4 2 3 1.70504 53.71568 6.76
2 4 4 1.69289 54.13229 6.06
4 2 4 1.68609 54.36876 7.08
1 4 8 1.68549 54.38977 27
1 3 13 1.68299 54.47701 18
4 1 8 1.67710 54.68430 8.24
4 2 5 1.66263 55.20069 10.9
0 3 14 1.65229 55.57582 7.61
4 2 6 1.63524 56.20624 83.7
4 1 9 1.63427 56.24283 25.9
0 4 10 1.63246 56.31056 10.1
4 0 10 1.62434 56.61754 14.7
4 1 10 1.59007 57.95189 5.46
4 2 8 1.57122 58.71473 30.8
3 3 10 1.56101 59.13673 8.17

Figure A.12. List of indexed hkl reflections obtained from XRD pattern of RbBa0.15Sr1.85Nb3O10
h k l d (Å) 2θ Intensity
0 0 2 15.18911 5.81387 888
0 2 0 3.88558 22.86871 165
0 0 8 3.79728 23.40796 95.4
2 1 1 3.46814 25.66566 118
1 2 4 3.16311 28.18939 93.8
1 1 8 3.12676 28.52399 143
1 0 9 3.09868 28.78811 92.5
0 1 9 3.09594 28.81414 422
2 0 6 3.09333 28.83892 1.22e+003
0 2 6 3.08247 28.94275 991
0 0 10 3.03782 29.37764 237
1 2 5 3.01924 29.56257 571
2 0 7 2.90381 30.76619 105
1 2 6 2.86747 31.16594 139
1 0 10 2.83142 31.57301 565
2 2 0 2.75521 32.47017 2.08e+003
2 0 8 2.72318 32.86276 248
2 1 7 2.72011 32.90089 82.9
2 2 9 2.13441 42.31036 86
2 0 12 2.12459 42.51529 97.5
3 0 9 2.06222 43.86676 280
2 3 6 1.98600 45.64322 110
3 0 10 1.97747 45.85151 116
1 0 15 1.96045 46.27242 102
2 2 11 1.95050 46.52235 340
4 0 1 1.94968 46.54321 848

229
0 4 0 1.94279 46.71789 74.2
1 3 10 1.91122 47.53663 93.1
1 1 15 1.90090 47.81089 141
1 4 2 1.87104 48.62262 217
2 2 12 1.86413 48.81482 105
2 4 3 1.71451 53.39529 121
3 0 14 1.66723 55.03545 87.7
1 2 16 1.66663 55.05682 121
4 1 9 1.65223 55.57819 85.9
4 2 6 1.65017 55.65345 233
1 0 18 1.64965 55.67255 318
1 4 9 1.64602 55.80593 219
2 4 6 1.64522 55.83554 112
0 4 10 1.63670 56.15166 170
2 2 15 1.63181 56.33521 106
2 0 17 1.62508 56.58942 100
4 2 7 1.61940 56.80572 216
4 1 10 1.60767 57.25861 122
4 2 8 1.58595 58.11673 85.4
2 3 13 1.58579 58.12328 89
2 4 8 1.58156 58.29379 125
3 3 10 1.57182 58.69020 90.8

Figure A.13. List of indexed hkl reflections obtained from XRD pattern of RbBa0.3Sr1.7Nb3O10
h k l d (Å) 2θ Intensity
0 0 2 15.13804 5.83350 15.3
2 0 1 3.96155 22.42444 10.1
0 2 0 3.93215 22.59432 6.55
2 1 2 3.46769 25.66899 11.7
2 1 3 3.35927 26.51232 5.97
2 1 4 3.22326 27.65284 31.3
2 0 6 3.13263 28.46949 8.09
0 1 9 3.09292 28.84282 18.6
2 1 5 3.07045 29.05851 10.2
1 2 5 3.04847 29.27275 35
0 0 10 3.02761 29.47899 49.2
2 0 7 2.93504 30.43088 10.8
1 1 9 2.88444 30.97790 7.01
1 0 10 2.83125 31.57497 47.8
2 2 0 2.80272 31.90491 22.6
2 2 1 2.79079 32.04499 44.1
2 2 2 2.75588 32.46198 9.46
0 3 1 2.61166 34.30856 6.45
1 3 8 2.08064 43.45861 5.9
0 3 9 2.06775 43.74345 21.4
2 3 5 2.06099 43.89424 19.2
2 1 12 2.05894 43.94038 8.37
4 0 1 1.99362 45.45916 6.61

230
4 0 2 1.98078 45.77054 34.6
0 4 0 1.96607 46.13248 11.9
0 4 2 1.94970 46.54263 5.32
3 3 1 1.86493 48.79237 11
4 2 1 1.77813 51.34233 6.82
2 3 10 1.77543 51.42614 5.86
1 3 12 1.77256 51.51568 7.41
2 4 5 1.69369 54.10474 8.79
0 0 18 1.68200 54.51165 8.27
2 1 16 1.67113 54.89597 7.9
4 0 10 1.66758 55.02295 5.39
1 2 16 1.66756 55.02349 5.24
2 4 6 1.66527 55.10569 6.48
2 3 12 1.65466 55.48938 8.28
0 4 10 1.64891 55.69976 6.58
2 2 15 1.63788 56.10775 7.02
2 4 7 1.63346 56.27316 5.58
3 3 9 1.63343 56.27422 5.6
2 0 17 1.62669 56.52811 8.44
4 0 11 1.61687 56.90280 12.7
3 0 15 1.60878 57.21541 5.42

Figure A.14. List of indexed hkl reflections obtained from XRD pattern of RbBa0.6Sr1.4Nb3O10
h k l d (Å) 2θ Intensity
0 0 2 15.36151 5.74857 387
2 0 0 3.88861 22.85070 132
2 0 1 3.85783 23.03549 502
0 0 8 3.84038 23.14160 97.2
0 1 7 3.83323 23.18534 71.4
0 2 2 3.81203 23.31609 121
1 1 6 3.75778 23.65754 431
2 0 3 3.63536 24.46635 296
2 1 0 3.48628 25.52984 398
1 0 8 3.44344 25.85293 174
1 2 2 3.42296 26.01030 166
2 1 2 3.39982 26.19045 109
1 2 3 3.32144 26.81992 82.7
2 0 5 3.28589 27.11559 108
1 1 8 3.15470 28.26614 123
0 1 9 3.13176 28.47752 106
0 2 6 3.12017 28.58558 96.1
2 0 6 3.09683 28.80566 195
0 0 10 3.07230 29.04064 669
1 2 5 3.04860 29.27144 1.98e+003
2 1 5 3.03222 29.43313 819
2 0 7 2.91057 30.69297 99
1 0 10 2.85742 31.27830 84.9
2 2 0 2.76596 32.34048 149
1 2 7 2.74181 32.63329 367
2 0 8 2.73245 32.74817 502

231
2 1 7 2.72987 32.77996 292
2 2 2 2.72218 32.87517 695
1 1 10 2.68588 33.33242 265
1 0 11 2.62863 34.08026 127
3 1 7 2.14742 42.04185 77.2
2 0 12 2.13838 42.22796 75.6
2 3 3 2.12734 42.45784 84.3
1 0 14 2.11203 42.78058 70.5
2 3 4 2.09251 43.19959 135
0 3 9 2.08011 43.47016 75.2
3 1 8 2.07281 43.63112 70.6
1 1 14 2.03986 44.37309 74.8
2 0 13 2.01958 44.84268 77.5
2 3 6 2.00171 45.26501 126
3 2 6 1.99397 45.45066 142
4 0 1 1.94042 46.77837 143
1 3 10 1.93248 46.98223 465
4 0 2 1.92891 47.07421 378
3 1 10 1.92136 47.27063 116
4 1 0 1.88756 48.17017 86.1
4 1 2 1.87347 48.55570 179
1 0 16 1.86421 48.81249 228
1 4 4 1.85123 49.17726 78.1
3 3 1 1.84066 49.47860 76.6
2 2 13 1.79677 50.77186 84.8
1 4 6 1.78747 51.05485 158
1 0 17 1.76033 51.89996 113
2 4 3 1.73038 52.86740 96.1
0 2 16 1.72570 53.02190 91.3
3 2 11 1.71105 53.51178 95.5
0 0 18 1.70683 53.65451 73.9
4 1 8 1.69400 54.09402 246
2 1 16 1.68194 54.51383 88.7
4 2 5 1.67697 54.68912 190
1 0 18 1.66716 55.03801 206
3 1 14 1.63827 56.09338 193
1 1 18 1.63097 56.36681 533
2 4 7 1.63005 56.40135 75
3 3 9 1.62240 56.69111 205
0 3 15 1.61443 56.99654 132
4 2 8 1.58728 58.06343 78.1
0 1 19 1.58392 58.19864 86.5
3 1 15 1.57463 58.57508 72
0 2 18 1.56588 58.93454 133

232
Appendix B: C1-6/CaxSr2-xNb3O10 TGA Data

C1/Sr2Nb3O10 TGA
100 0.16
0.14
99.5

Derivative Weight %
0.12
0.1
Weight %

99
0.08
0.06
98.5
0.04

98 0.02
0
97.5 -0.02
100 200 300 400 500 600 700
Temperature, Degrees C

Figure B.1. TGA data for C1/Sr2Nb3O10 with temperatures ranging from 150 to 800 °C. See Table B.1
for mass loss % and coverage %.

C3/Sr2Nb3O10
101 0.07

100 0.06

99 0.05

98 0.04

97 0.03

96 0.02

95 0.01

94 0

93 -0.01
0 100 200 300 400 500 600 700 800

Figure B.2. TGA data for C3/Sr2Nb3O10 with temperatures ranging from 150 to 800 °C. See Table B.1
for mass loss % and coverage %.

233
C6/Sr2Nb3O10 TGA
100 0.16
99 0.14

Derivative Weight %
98 0.12
0.1
Weight %

97
0.08
96
0.06
95
0.04
94 0.02
93 0
92 -0.02
100 200 300 400 500 600 700
Temperature, Degrees C

Figure B.3. TGA data for C6/Sr2Nb3O10 with temperatures ranging from 150 to 800 °C. See Table B.1
for mass loss % and coverage %.

C1/Ca0.4Sr1.6Nb3O10 TGA
100 0.16
0.14
99.5 Derivative Weight %
0.12
99 0.1
Weight %

0.08
98.5
0.06
98 0.04
0.02
97.5
0
97 -0.02
100 200 300 400 500 600 700
Temperature, Degrees C

Figure B.4. TGA data for C1/Ca0.4Sr1.6Nb3O10 with temperatures ranging from 150 to 800 °C. See Table
B.1 for mass loss % and coverage %.

234
C3/Ca0.4Sr1.6Nb3O10
101 0.05

100 0.04
99
0.03
98
0.02
97
0.01
96
0
95

94 -0.01

93 -0.02
0 100 200 300 400 500 600 700 800 900

Figure B.5. TGA data for C3/Ca0.4Sr1.6Nb3O10 with temperatures ranging from 150 to 800 °C. See Table
B.1 for mass loss % and coverage %.

C6/Ca0.4Sr1.6Nb3O10
101 0.16
100 0.14
99
0.12
98
0.1
97
96 0.08
95 0.06
94
0.04
93
0.02
92
91 0
90 -0.02
0 100 200 300 400 500 600 700 800 900

Figure B.6. TGA data for C6/Ca0.4Sr1.6Nb3O10 with temperatures ranging from 150 to 800 °C. See Table
B.1 for mass loss % and coverage %.

235
C1/Ca1Sr1Nb3O10
101 0.16
0.14
100
0.12
99
0.1
98 0.08

97 0.06
0.04
96
0.02
95
0
94 -0.02
0 100 200 300 400 500 600 700 800 900

Figure B.7. TGA data for C1/Ca1Sr1Nb3O10 with temperatures ranging from 150 to 800 °C. See Table
B.1 for mass loss % and coverage %.

C3/Ca1Sr1Nb3O10
101 0.25
100
0.2
99
98 0.15
97
0.1
96
95 0.05
94
0
93
92 -0.05
0 100 200 300 400 500 600 700 800 900

Figure B.8. TGA data for C3/Ca1Sr1Nb3O10 with temperatures ranging from 150 to 800 °C. See Table
B.1 for mass loss % and coverage %.

236
C6/Ca1Sr1Nb3O10
102 0.25

100
0.2
98
0.15
96

94 0.1

92
0.05
90
0
88

86 -0.05
0 100 200 300 400 500 600 700 800 900

Figure B.9. TGA data for C6/Ca1Sr1Nb3O10 with temperatures ranging from 150 to 800 °C. See Table
B.1 for mass loss % and coverage %.

C1/Ca1.5Sr0.5Nb3O10
101 0.1

100 0.08

99 0.06

98 0.04

97 0.02

96 0

95 -0.02
0 100 200 300 400 500 600 700 800 900

Figure B.10. TGA data for C1/Ca1.5Sr0.5Nb3O10 with temperatures ranging from 150 to 800 °C. See
Table B.1 for mass loss % and coverage %.

237
C3/Ca1.5Sr0.5Nb3O10
101 0.07

100 0.06

99 0.05

98 0.04

97 0.03

96 0.02

95 0.01

94 0

93 -0.01
0 100 200 300 400 500 600 700 800

Figure B.11. TGA data for C3/Ca1.5Sr0.5Nb3O10 with temperatures ranging from 150 to 800 °C. See
Table B.1 for mass loss % and coverage %.

C6/Ca1.5Sr0.5Nb3O10
101 0.16
100 0.14
99
0.12
98
0.1
97
96 0.08
95 0.06
94
0.04
93
0.02
92
91 0
90 -0.02
0 100 200 300 400 500 600 700 800 900

Figure B.12. TGA data for C6/Ca1.5Sr0.5Nb3O10 with temperatures ranging from 150 to 800 °C. See
Table B.1 for mass loss % and coverage %.

238
C1/Ca1.8Sr0.2Nb3O10 TGA
100 0.16
0.14

Derivative Weight %
99.5 0.12
0.1
Weight %

99
0.08
0.06
98.5
0.04

98 0.02
0
97.5 -0.02
100 200 300 400 500 600 700
Temperature, Degrees C

Figure B.13. TGA data for C1/Ca1.8Sr0.2Nb3O10 with temperatures ranging from 150 to 800 °C. See
Table B.1 for mass loss % and coverage %.

C3/Ca1.8Sr0.2Nb3O10
101 0.035
0.03
100
0.025
99 0.02
0.015
98
0.01
97
0.005

96 0
-0.005
95
-0.01
94 -0.015
0 100 200 300 400 500 600 700 800 900

Figure B.14. TGA data for C3/Ca1.8Sr0.2Nb3O10 with temperatures ranging from 150 to 800 °C. See
Table B.1 for mass loss % and coverage %.

239
C6/Ca1.8Sr0.2Nb3O10
102 0.25

100
0.2
98
0.15
96

94 0.1

92
0.05
90
0
88

86 -0.05
0 100 200 300 400 500 600 700 800 900

Figure B.15. TGA data for C6/Ca1.8Sr0.2Nb3O10 with temperatures ranging from 150 to 800 °C. See
Table B.1 for mass loss % and coverage %.

C1/Ca2Nb3O10 Microwave Grafting TGA


100 0.05

0.04
Derivative Weight %

99 0.03
Weight %

0.02

98 0.01

97 -0.01
100 200 300 400 500 600 700
Temperature, Degrees C

Figure B.16. TGA data for C6/Ca2Nb3O10 with temperatures ranging from 150 to 800 °C. See Table
B.1 for mass loss % and coverage %.

240

You might also like