Sei sulla pagina 1di 199

Development o f 2-D Nanostructured Layered Hydroxy Salts (LHSs) and

Hydroxy Double Salts (HDSs) for New Applications: Anionic Exchange


Kinetics and Polymer Modification

by

Everson Kandare

BSc (Hons) Chemistry, University o f Zimbabwe, 2000

A Dissertation Submitted to the Faculty o f the


Graduate School, Marquette University,
In Partial Fulfillment of
The Requirements for
The Degree o f
Doctor o f Philosophy

Milwaukee, Wisconsin

April, 2006

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

UMI Number: 3210968

Copyright 2006 by
Kandare, Everson

All rights reserved.

INFORMATION TO USERS

The quality of this reproduction is dependent upon the quality of the copy
submitted. Broken or indistinct print, colored or poor quality illustrations and
photographs, print bleed-through, substandard margins, and improper
alignment can adversely affect reproduction.
In the unlikely event that the author did not send a complete manuscript
and there are missing pages, these will be noted. Also, if unauthorized
copyright material had to be removed, a note will indicate the deletion.

UMI
UMI Microform 3210968
Copyright 2006 by ProQuest Information and Learning Company.
All rights reserved. This microform edition is protected against
unauthorized copying under Title 17, United States Code.

ProQuest Information and Learning Company


300 North Zeeb Road
P.O. Box 1346
Ann Arbor, Ml 48106-1346

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Everson Kandare
Jeanne M. Hossenlopp
Abstract
Development o f 2-D Nanostructured Layered Hydroxy Salts (LHSs) and Hydroxy
Double Salts (HDSs) for New Applications: Anionic Exchange Kinetics and Polymer
Modification

Inorganic/organic 2-D nanostructured hybrid materials with tunable interlayer


spacing have been prepared and characterized via X-ray diffraction (XRD), Fourier
transform infrared (FTIR), diffuse reflectance spectroscopy (DRS), nuclear magnetic
resonance
spectroscopy (NMR), transmission electron microscopy (TEM),
thermogravimetric analysis (TGA), and TGA-FTIR for various applications such as
catalysis, anion exchange, fire retardancy, and controlled release o f ions. The kinetics
and mechanism o f the anionic exchange reactions were investigated and a correlation
between structure and reactivity o f model layered hydroxy double salts was attempted.
Combined analysis o f the evolution o f solid state products via XRD and solution phase
anion concentrations via NMR provided the data necessary to evaluate kinetic models
and fully characterize the reactivity o f these compounds. Exchanging acetate anions with
other anions o f varying anionic chain lengths provided an insight into different
mechanisms via which these reactions can be accomplished. The anionic exchange
mechanisms were found to be dependent on the diffusion o f anions in and out o f the
interlayer space as well as on nucleation growth in accordance with the Avrami-Erofeev
geometric model. In the long run, the anionic release profiles may be used to model the
release o f drugs from nanostructured drug delivery systems. The thermal degradation of
acetate containing layered compounds was studied and the results obtained suggest that
the presence o f polycrystalline ZnO in the residue promotes the ketonization channel o f
acetic acid into acetone. An understanding o f the degradation mechanism o f the HDSs
and LHSs by themselves afforded us to explore the role o f these layered additives in
enhancing polymer fire retardancy. For example, from cone calorimetry measurements,
polymer nanocomposites containing modified hydroxy double salts (HDSs) and layered
hydroxy salts (LHSs) clays have been shown to exhibit lower total heat release when
compared with untreated virgin polymer. The presence o f certain metal ions in their low
valent oxidation states may promote chemical (redox) effects that may serve to slow
down or prevent the depolymerization process. Some metallic species may promote char
formation, which prevents mass and heat transfer hence slows down the burning process.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Preface
Hydroxy double salts (HDSs) and layered hydroxy salts (LHSs) are a class of 2-D
nanostructured materials with nanometer sized galleries. These compounds exhibit
interesting physical properties and have many potential and demonstrated applications in
various fields, including: anion separations, drug delivery, environmental water
decontamination, catalysis, magnetism, and fire retardancy. These compounds have
structural properties that can be utilized as design parameters to fine-tune for improved
effectiveness in chosen areas of application.
The goal of this work is to explore the various synthetic methods and
characterization techniques for HDSs and LHSs that will yield important structural
information and corresponding reactivities of these nanomaterials. With this fundamental
insight into structure/reactivity relationships, these compounds can be tailored for
specific applications as detailed above.
Chapter 1 gives a general introduction of these nanostructured materials
and provides a detailed literature review on structurally similar smectite clays. In Chapter
2, preparation and characterization o f three model compounds, zinc copper acetate
(ZCA), zinc nickel acetate (ZNA), and zinc hydroxy acetate (ZHA) are described.
Anionic exchange kinetics were performed using the three model compounds and the
results were correlated to their structures and the exchange anion size. The effect of
varying the exchange reaction conditions such as temperature, time, and anion
concentration on the morphology of the exchange products are discussed. Chapter 3
presents an investigation of the thermal degradation of ZCA, ZNA, and ZHA and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

highlights the possible catalytic effect that ZnO may have on the ketonization of acetic
acid to acetone in the gas phase. The presence o f a second metal with Zn is investigated
to see if that has an effect on the ZnO crystal growth. Chapter 4 describes the efforts
taken to improve the thermal stability and fire retardancy of poly (methyl methacrylate)
using copper-containing layered materials. Thermal stability and flammability are
evaluated using thermogravimetric analysis and cone calorimetry respectively.
Chapter 5 demonstrates the use o f the Flynn-Wall-Ozawa (FWO) isoconversional
method to estimate the apparent activation energy as a function o f conversion fraction for
pure polystyrene versus its respective composites containing putative thermal stabilizers.
Copper hydroxy dodecyl sulfate is used as a potential fire retardant with poly (vinyl
esters) and this work is presented in Chapter 6. The summary of this work and suggested
future endeavors are described in Chapter 7.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Acknowledgements
My sincere gratitude goes to my research advisor, Prof. Jeanne M. Hossenlopp for
her unwavering support throughout this research project. Her scientific guidance is
greatly appreciated. I would also like to thank Prof. Daniel Haworth, Prof. Chieu Tran,
and Prof. Scott Reid for affording time out their schedules to serve on my research
committee. I am also thankful to the Marquette University Committee on Research and
the Office of Naval Research for providing the financial support for this work, Prof.
Daniel Sem for help with optimization of 'H NMR acquisition parameters, Prof. Charles
Wilkie and coworkers for the fruitful discussions in polymer modification, especially Dr.
Grace Chigwada, and Prof Joseph Collins for XRD experiments. I would also like to
thank former group members, Dr. Yolanda Jones and Dr. Hongmei Deng, undergraduate
coworkers, Karen Rader, Elizabeth Vissat, Danielle Brusse, and Daniel Hall (Alcorn
State University) particularly for their work in the exploration of new synthesis methods
for layered hydroxy salts. I am grateful to have a loving family constantly showering
encouragements. My dearest Gleny, you are a pillar of support and cherished you will
stay.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Dedication
In loving memory o f my late mother
Ettah Dhlodhlo-Kandare

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Tables o f Contents
List of Figures........................................................................................................................viii
List of Tables

............................................................................................................xii

Chapter 1

2-Dimensional Nanostructured Materials and their Applications................. 1

1. 1

Introduction....................................................................................................... 1

1.1.1

Cationic Clays...................................................

1.1.2

Anionic Clays....................................................................................................5

1.1.3

Assembly of Brucite-like Anionic Clays....................................................... 7

1.1.4

Motivation for the study................................................................................. 11

Chapter 2

Hydroxy Double Salt Anion Exchange Kinetics and Thermodynamics:

Effects of Precursor Structure and Anion Size........................................... 14


2. 1

Introduction.................................................................................................... 14

2.1.1

Anionic Exchange Kinetics........................................................................... 14

2.1.2

Thermodynamics of Anionic Exchange andChain Packing M odes

2.2

Experimental.................................................................................................. 18

2.2.1

Anionic Exchange Kinetics........................................................................... 18

2.2.2

Thermodynamics of Anionic Exchange andChain Packing M odes

2.3

Results and Discussion.................................................................................. 23

2.3.1

Anionic Exchange Kinetics........................................................................... 23

2.3.1.1

Conclusions.................................................................................................... 43

2.3.2

Thermodynamics of Anionic Exchange andChain Packing M odes

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

17

22

44

2.3.2.1

Conclusions.................................................................................................... 57

Chapter 3

Thermal Degradation o f Acetate Intercalated Hydroxy Double Salts


(HDSs) and Layered Hydroxy Salts (LHSs).............................................. 59

3.1

Introduction.................................................................................................... 59

3.2

Experimental................................................................................................ 60
61

3.3

Results and Discussion.....................

3.4

Conclusions.................................................................................................... 80

Chapter 4

Thermal Stability of Hydroxy Double Salt (HDS) and LayeredHydroxy


Salt (LHS)/Poly(methyl methacrylate) (PMMA) Composites.................82

4. 1

Introduction................................................................................................ 82

4.2

Experimental................................................................................................83

4.3

Results and Discussion......................................................

86

4.4....................... Conclusions...............................................................................................102

Chapter 5

Thermal Degradation Kinetics o f Polystyrene (PS) Nanocomposites.. 104

5.1

Introduction............................................................................................... 104

5.2

Experiment................................................................................................105

5.3
5.4

Results and Discussion

......................................................................... 107

Conclusions............................................................................................... 124

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Chapter 6

Nanostractured Layered Copper Hydroxy Dodecyl Sulfate: A Potential


Fire Retardant for Poly(vinyl ester) (PVE)..............................................126

6.1

Introduction................................................................................................. 126

6.2

Experimental............................................................................................... 128

6.3

Results and Discussion.............................................................................. 129

6.4

Conclusions..................................................................................

Chapter 7

Summary and Future Work........................................................................ 147

7.1

Summary......................................................................................................147

7.2

Proposed Future Experiments....................................................................150

7.2.1

Anionic Exchange Kinetics......................................................................... 150

7.2.2

In-situ Anion Release Profile using Ultraviolet-Visible (UV-VIS)

7.2.3

Polymer Modification..................................................................................155

7.2.4

Predictive Modeling of Thermal Degradation of Polymers.....................156

145

154

References............................................................................................................................. 164
Appendix A ........................................................................................................................... 179
Appendix B ........................................................................................................................... 180
Appendix C ............................................................................................................................181
Appendix D ........................................................................................................................... 182

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

List o f Figures
Figure 1.1.2.1 Schematic representation o f the layered double hydroxide (LDH)............. 5
Figure 1.1.3.1 The idealized structure of an HDS................................................................. 8
Figure 1.1.3.2 A schematic representation o f the structure of a layered hydroxy salt....... 9
Figure 2.1.1.1 First-order and second order staging in anionic exchange reactions......... 16
Figure 2.3.1.1 PXRD data for ZNA HDS and ZHA.......................................................... 24
Figure 2.3.1.2 PXRD data for ZCA HDS............................................................................ 25
Figure 2.3.1.3 Diffuse reflectance spectra of ZCA, zinc copper butyrate (ZCB), zinc
copper hydroxycinnamate (ZCH), and zinc copper glutamate (ZCG) HDS...................... 27
Figure 2.3.1.4 FTIR spectra of layered compounds and reference metal acetate
precursors................................................................................................................................ 28
Figure 2.3.1.5 Interlayer spacings for Zn/Cu HDS n-alkyl homologous series................. 30
Figure 2.3.1.6 H1NMR spectrum of the filtrate obtained after 7200 seconds exchanging
the acetate anions with octanoate anions in ZCA.................................................................31
Figure 2.3.1.7 Extent of the reaction data for release of acetate and corresponding loss of
octanoate from solution obtained for reaction with ZCA HDS...........................................33
Figure 2.3.1.8 Corresponding Sharp-Hancock plot for release of acetate into solution... 34
Figure 2.3.1.9 Acetate solution phase concentrations as a function of time following
exposure of ZCA HDS to butyrate and octanoate anions.................................................... 35
Figure 2.3.1.10 PXRD data for solid collected at times (s) indicated following exposure
of ZCA to each of the three anions........................................................................................37
Figure 2.3.1.11 PXRD trace for ZCB....................................................................................39
Figure 2.3.1.12 Extent o f reaction as a function of time for acetate anions released from

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

ZHA exposed to octanoate solution and corresponding Sharp-Hancock plot (inset)........40


Figure 2.3.1.13 Extent of reaction as a function of time for acetate anions released from
ZNA HDS................................................................................................................................41
Figure 2.3.1.14 PXRD data obtained for solid materials recovered from the reaction of
ZNA HDS with octanoate...................................................................................................... 42
Figure 2.3.2.1 PXRD patterns following exchange o f acetate by 4-hydroxycinnamate... 44
Figure 2.3.2.2 Proposed model structures for the bilayer phase and the interpenetrating
monolayer phase of the exchange products.......................................................................... 46
Figure 2.3.2.3 PXRD patterns following exchange of acetate by 4-hydroxycinnamate... 47
Figure 2.3.2.4 PXRD patterns following exchange of acetate by 4-hydroxycinnamate... 49
Figure 2.3.2.5 PXRD patterns following exchange o f acetate by 4-hydroxycinnamate... 51
Figure 2.3.2.6 FTIR spectra of 4-hydroxycinnamate, exchanged products, and ZCA......53
Figure 23.2.1 Thermogravimetric analysis curves for the exchanged products................54
Figure 3.3.1

TGA and corresponding DTG and DTA curves of ZHA in air.................. 62

Figure 3.3.2 FTIR spectra of evolved gaseous products of ZHA in air..............................63


Figure 3.3.3 FTIR spectra of evolved gaseous products collected at 328 C during thermal
combustion of ZHA in air...................................................................................................... 64
Figure 3.3.4 XRD analysis of residual products of ZHA in air...........................................66
Figure 3.3.5 FTIR spectra of residual products collected after heating ZHA in air

68

Figure 3.3.6 TGA and corresponding DTG and DTA curves o f ZNA in air..................... 70
Figure 3.3.7 FTIR spectra of evolved gaseous products of ZNA in air..............................71
Figure 3.3.8 XRD analysis of the inorganic residue from heating ZNA in air.................. 72
Figure 3.3.9 FTIR spectra of residual products collected after heating ZNA.................... 74

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Figure 3.3.10 TGA and corresponding DTG curves of ZCA in air.................................... 75


Figure 3.3.11 FTIR spectra of evolved gaseous products of ZCA in air............................ 76
Figure 3.3.12 XRD analysis of the residue from thermal combustion of ZCA in air

77

Figure 3.3.13 FTIR spectra of residual products collected after heating ZCA.................. 78
Figure 4.3.1 PXRD pattern for Cu2 (0 H)3N 0 3 , (gerhardtite) and CH M ............................ 87
Figure 4.3.2 FTIR spectral patterns of CHN and the exchange product CHM.................. 87
Figure 4.3.3 PXRD data for ZCA HDS and ZCM............................................................... 89
Figure 4.3.4 XRD patterns of PMMA-ZCM2% and PMMA CHM2%.............................89
Figure 4.3.5 TEM image for PMMA-CHM2% composites (bulk polymerization)...........91
Figure 4.3.6 TEM image for PMMA-ZCM2% composites (bulk polymerization)...........92
Figure 4.3.7 TGA curves for ZCM, pure PMMA, PMMA-ZCM2%, and PMMA-ZCM4%
composites............................................................................................................................... 92
Figure 4.3.8 TGA curves for CHM, pure PMMA, PMMA-CHM2%, and PMMA
CHM4% composites...............................................................................................................93
Figure 4.3.9 TGA and DTA curves for pure PMMA, and PMMA-ZCM2%.....................96
Figure 4.3.10 TGA and DTA curves for pure PMMA and PMMA-CHM2%................... 97
Figure 4.3.11 XRD patterns of PMMA-CHM residue from cone calorimetry.................. 98
Figure 4.3.12 XRD patterns of PMMA-ZCM residue from cone calorimetry.................. 99
Figure 4.3.13 Heat release rate curves for PMMA-ZCM2%, PMMA-CHM2%, and pure
PMMA................................................................................................................................... 101
Figure 4.3.14 Heat release rate curves for PMMA-ZCM4%, PMMA-CHM4%, and pure
PMMA................................................................................................................................... 101
Figure 5.3.1 Structures of the salts used to prepare the organically modified clays

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

107

Figure 5.3.2 XRD data for Cloisite 10A, VB16 clays, and their PS/clay composites.... 108
Figure 5.3.3 TEM image for PS-10A.................................................................................110
Figure 5.3.4 TEM image for PS-VB16.............................................................................. 110
Figure 5.3.5 TGA and DTG curves for pure PS, PS-VB16, PS-PYC16, PS-10A, and PS
QC16.................................................... .,................................................................................ 112
Figure 5.3.6 Curves of mass loss differences for PS composites....................................114
Figure 5.3.7 Plot of Ea versus extent of decomposition, a for PS and its composites.... 118
Figure 5.3.8 Heat release rate curves pure PS and its composites.................................... 121
Figure 6.3.1 Chemical structures of bisphenol-A and novalac epoxy vinyl resins

130

Figure 6.3.2 XRD data for partially exchanged CHDS and PVE-n composites at n= 1,3,
5, 7, and 10% loadings......................................................

131

Figure 6.3.3 Schematic representations of the layered copper hydroxy nitrate (CHN).. 132
Figure 6.3.4 TEM images for PVE/CHDS-5...................................................................... 134
Figure 6.3.5 TGA and DTG curves for pure PVE and its composites..............................135
Figure 6.3.6 Curves o f mass loss differences for PVE composites.................................. 137
Figure 6.3.7 HRR curves for PVE and its composites from cone calorimetry................ 139
Figure 6.3.8 Percent reduction in total heat release (THR) vs. % additive; observed
reductions and calculated reductions.................................................................................. 141
Figure 6.3.9 XRD pattern of residue after heating PVE/CHDS-10 to indicated
temperatures in TGA and residue from cone calorimetry experiments.......................... 142
Figure 6.3.10 FTIR traces for pure PVE, PVE/CHDS-10, and TGA residue of
PVE/CHDS-10 heated to different temperatures indicated in the plot............................. 144

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

x ii

Figure 7.4.2.1 TGA curves for pure PE and PE-ZCHS-5.................................................. 158


Figure 7.4.2.2 Mass loss difference curve for PE-ZCHS-5...............................................159
Figure 7.4.2. 3 Experimental and reconstructed a-T curves for PE and PE-ZCHS-5 .... 163

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

List o f Tables
Table 2.3.1.1 Carboxylate vibrational frequencies in the precursor metal acetates .........29
Table 2.3.1.2 Kinetic data for exchange reactions............................................................... 36
Table 3.3.1 ZHA Thermal residue: relative intensities o f ZnO XRD peaks...................... 66
Table 3.3.2 Average crystallite sizes of metal oxide phases................................................67
Table 4.3.1TGA data for pure PMMA and its composites in air and N 2 ............................94
Table 4.3.2 Cone data for pure PMMA and its composites......................................... 100
Table 5.3.1 TGA data for polystyrene nanocomposites...............................................114
Table 5.2.2 Cone calorimetry data for polystyrene composites at a flux of 35 kW/m2.. 123
Table 6.3.1 TGA data for PVE composites...........................

136

Table 6.3.2 Cone calorimetry data for PVE composites at 35 kW/m2.......................140


Table 7.4.2.1 TGA data for ZCHS, poly(ethylene) and PE-ZCHS-5............................... 158

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Chapter 1

1.1

2-Dimensional Nanostructured Materials and their Applications

Introduction
Layered inorganic or inorganic/organic hybrid compounds with nanodimensional

interlayer spacings have been developed for applications such as catalysts,12 catalyst
1

precursors, ' catalyst support material, adsorbents, antacids, ion exchangers, water
purification,7 and fire retardancy.10' 15 For example, polymer nanocomposites containing
modified cationic (smectite) clays have been shown to exhibit improved mechanical
properties as well as lowered peak heat release when compared with untreated virgin
polymer.16' 19
Layered double hydroxides (LDHs) are a subgroup of 2-D nanostructured
materials. LDHs, with a generic formula [M2+i.xM3+x(OH)2]x+Axn-zH20, consist of
positively charged metal hydroxide layers containing divalent and trivalent metal ions
and inorganic or organic anions, An', which serve to balance the charge as well as control
the interlayer spacing. LDHs can be considered complementary to smectite clays and the
ability to exchange intercalated ions has been a key factor in optimizing these materials
for different applications. Similarly, anion exchange reactions have been utilized with
basic metal hydroxides such as intercalated copper hydroxide, Cu2 (OH)3 X, where X is an
intercalated anion.20'22 The magnetic properties o f intercalated copper hydroxides have
been shown to exhibit a striking dependence on the interlayer spacing.20'22
Hydroxy double salts, (HDSs), are a more recently emerging class o f compounds
that hold similar promise for development as new nanostructured materials for a wide
range of applications. HDSs, [(M2+i.xMe2+i+x)(0H)3(i.y)]+An'(i+3y)/n-zH20, are similar to

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

LDHs except that the metal hydroxide layer contains two divalent metals, designated here
as M2+ and Me2+. These metal hydroxide sheets have a partial positive charge. Charge
balancing anions are thus sandwiched between hydroxylated metal sheets. Weak
hydrogen bonding, van der W aals, and electrostatic interactions between the metal
hydroxide sheets and the anions are responsible for holding the structures together. What
is most exciting about these materials is the availability o f a wide series of both intralayer
cations and the anions that can occupy the interlayer spacing. Cations are selected on
basis of their charge and size while the anions can vary from simple straight chain
carboxylates to complex polyoxometalates, (POMs).23 These materials exhibit anionic
exchange properties allowing rapid diffusion of anions from exchange solutions into
gallery spaces replacing the anions originally contained therein. This unique ability
allows HDSs to be used as precursors for the synthesis of new-layered nanostructured
materials through anion exchange.
Systematic incorporation of different metal ions and anions in these materials
results in structural variations that are expected to show different reactivities. Morioka
and coworkers7 have reviewed the synthesis, preparative methods, and post synthesis
modification of these layered materials. They extensively discussed how 2-D layered
materials are made via co-precipitation (direct synthesis method), ion-exchange methods,
and calcination methods (thermal decomposition). The first two methods stated above are
easy to envisage, while the last method relies on the chemical "memory effect" possessed
by calcined products of these layered materials. Calcination o f the layered hydroxides
yields metal oxides that can then be rehydrated in a solution containing the target
interlayer anion to form lamellar structures again.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

The insertion of guest species into the solid matrix of the layered material is
known as an intercalation process. In this context, intercalation can thus be viewed as a
simplistic description of topotactic solid-state reactions in reference to insertion of guest
and the loss of incumbent anions.24 Topotacticity here is defined as the anionic exchange
process maintaining the structure o f the intralayer metal hydroxide sheets. Intercalation
reactions involving these layered materials are a result of the flexible layered structures,
which can liberally adjust to contain inserted guest anions of varying geometry. This
phenomenon has been recognized in natural mineral clays where the interlayer cations are
weakly bound as compared to intralayer cations, and thus can easily be exchanged with
other different cations from solution.25
Due to the structural similarity of LDHs and HDSs to cationic clays, we predict
similar reactivity as has been demonstrated with the latter. Adapting known cationic
exchange reactions to their anionic counterparts will open new opportunities for tailored
fabrication of 2-D nanostructured materials with a range of desirable properties. As a
result, a large group of compounds with a variety of chemical and physical properties can
be synthesized for specific applications. With this in mind, a brief background
concerning the chemistry of the cationic clays is therefore a necessity and is presented
below.

1.1.1 Cationic Clays


Cationic clays, also known as smectite clays, consist of stacked negatively
charged aluminosilicate or magnesiosilicate layers with interlayer spaces occupied by
positively charged alkali metal ions. These materials are derived from charge neutral

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

layers of pyrophyllite, AUSigC^oiPH^ and talc, Mg6Si80 2o(OH)4 by isomorphous


substitution o f Si(IV) by M(III) ions and Mg(II) by M(I) ions respectively. Si atoms
occupy tetrahedral sites on either sides of an octahedral four-atom thick close packing of
oxide and hydroxide ions median sheet. This structure is normally referred to as the
tetrahedral-octahedral-tetrahedral (T-O-T) packing geometry. Mg occupies all octahedral
lattice sites in talc while A1 occupies only two thirds of the octahedral sites in
pyrophyllite.26
Substitution of tetrahedral Si(IV) by M(III) in pyrophyllite and talc results in
negatively charged layered compounds of the form [AU(Si8 -xMinx)0 2 o(OH)4 ]x' and
[Mg6 (Si8 -xMinx)C)2 o(OH)4 ]x' respectively. On the other hand, partial substitution of
octahedral Al(III) by M(II) ions in pyrophyllite or Mg(II) by M(I) in talc yields [AI4.
x(Si8Mnx) 0 2o(OH)4]x' and [Mg6-x(Si8MIx) 0 2o(OH)4]x' respectively. A commonly known
cationic clay is sodium montmorillonite, a hydrated sodium calcium aluminum silicate
(Na, Ca) (Al, Mg)6 (Si4 0 io)3 (0H )6-zH20 . Montmorillonite has a 2:1 expanding crystal
lattice that allows water movement between the sheets, resulting in the swelling hence
reversible cationic exchange. Hectorite is another member of smectite clays and is
primarily composed of sodium magnesium lithium silicate
(Mg,Li)3 Si4 0 io(OH)2 Nao.3 -4 H 20 and has extremely good rheological properties i.e. as a
thickener in the oil drilling industry because of its swelling abilities in water.1 Most
cationic clays exhibit a range o f useful properties including among others, surface acidity,
and cationic diffusion that renders them useful as base materials for medicines,
cosmetics, catalysts, ion exchangers, corrosion protectors,27 and recently as additive for
potentially enhancing fire retardancy.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

1.1.2

Anionic Clays

Natures laboratory has been successful in the formulation o f cationic clays with
interesting properties hence our desire and motivation to design anionic clays, a
complementary class. Here, the metal hydroxide sheets are positively charged with
charge compensating anions occupying the gallery space together with water.
Hydrotalcite-like layered double hydroxides, (LDHs) anionic clays have been extensively
studied. ' These materials have been termed hydrotalcite-like in reference to their
isomorphism with the natural mineral hydrotalcite, (Mg 6 Al 2 ( 0 H )i 6 )(C 0 3 ) -4H20 . The
general formula for LDHs is [M2+i_xM3+x(OH)2]x+ An-zH20 , where M2+ and M3+ occupy
the octahedral lattice sites o f the brucite- (Mg(OH)2) like layers forming positively
charged hydroxide sheets, with anions, An"and H20 occupying the interlayer space.
Figure 1.1.2.1 shows a schematic representation o f the LDH structure.

Figure 1.1.2.1 Schematic representation o f the layered double hydroxide (LDH). M(OH)6
octahedral cells are joined together through edge sharing to form infinitely large
positively charged layers (in green), with the anions "sandwiched" between them (in
blue).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

The combination of the layered metal hydroxide sheets and intercalated anions
results in interesting 2-D inorganic/organic hybrids that contain nanometric domains. The
weak hydrogen bonding, van der Waals, and electrostatic forces that describe the
interaction between the anions and the positive metal hydroxide sheets dictates the
reactivity of these materials. Constantino and Pinnavaia28 have investigated the basic
properties exhibited by these compounds while varying the identity of the exchangeable
anions. Intercalation by long chains in these host materials results in the expansion of the
c-parameter (interlayer spacing) while maintaining the structural integrity o f the
intralayer structure. This is another way of utilizing topotactic exchange chemistry of
these materials to tailor anionic clays that possess desirable properties for targeted
applications. LDHs have been used in various aspects of chemistry including
pharmaceutical, catalysis, fire retardancy,31'34 ion mobility, and environment protection.35
Other classes of layered anionic clays are hydroxy double salts, (HDSs) and
layered hydroxy salts, (LHSs). Similar in structure to LDHs, these compounds however
contain divalent metal ions only. In this respect, they also possess the anionic exchange
abilities as shown for LDHs. Varying the metal identity and composition in these
compounds is expected to generate a wide range of chemical applications as seen for both
cationic and LDH clays. In this work, we set to explore the various methods by which
this novel class of 2-D layered materials can be synthesized, characterized, and applied to
the field of science.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

1.1.3

Assembly of Brucite-like Anionic Clays.

The framework of anionic clays is generally consistent with the brucite-like


(Mg(OH)i) structure. In brucite each Mg2+ ion is octahedrally coordinated to six OH'
anions and the resultant M(OH)6 octahedra units share their edges in order to build the
hydroxylated metal sheets. These metal hydroxide layers will in turn stack on top of each
other and are held together by weak electrostatic interactions. In LDHs, isomorphic
substitution of some of the Mg2+ ions by cations of a higher charge but similar radius,
such as Al3+, results in the brucite like sheets being positively charged. To maintain
electrical neutrality anions occupy the interlayer domains as shown in Figure 1.1.2.1
above. The LDHs have a wide selection of M2+/M3+ ratios and can accommodate a
variety of anions.
There are several methods by which positively charged metal hydroxide sheets
can be generated in these brucite-like structures. A number of possible synthetic
pathways have been investigated showing the steps that need to be followed to achieve
the above-mentioned objective. Rationalized below are some of the ways the
hydroxylated metal sheets may acquire a net partial positive charge.

a) Modification o f the intralayer structure


(1)

A partial positive charge is introduced into the hydroxylated metal sheets of

these anionic clays if there are vacant sites in the octahedrally coordinated brucite-like
structure. Insertion of twice as many divalent cations in the tetrahedral positions above
and below these vacant sites results in an increase in the cation/anion ratio. These
divalent cation enriched hydroxylated layers yield compounds of the formula, [(M2+i.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

x,Me2 +i+x)(OH)3 (i-y)]+An'(i+3 y)/n,zH2 0 . Zn2+ enriched zinc nickel acetate, (ZNA) HDSs
have been synthesized where the Zn2+ ions occupy tetrahedral positions while Ni2+ ions
are found in the octahedral environments9 , 3 6 , 37 as illustrated in Figure 1.1.3.1.

Figure 1.1.3.1 The idealized structure of an HDS with octahedrally-coordinated Me2+ in


grey, tetrahedrally-coordinated M2+ in black (above and below Me2+ vacancies), with
acetate anions bound to the M2+. Water is also contained in the interlayer space.

(2)

A class of compounds similar to that descried above is obtained when the

tetrahedral and octahedral divalent metal ions are the same. Here, the tetrahedrally
positioned cations are coordinated to three OH' ions forming part o f the intralayer
structure with the interlayer filling anion completing the geometry. Stahlin and Oswald38
reported synthesis and characterization o f the crystal structure o f zinc hydroxide nitrate,
Zn5 (0 H)8 (N 0 3 )2 -2 H 20 , formulated as [Zncta3 Zntetra2 (0 H)8 ][N 0 3 ]2 -2 H2 0 where the
tetrahedrally positioned Zn2+ ions are coordinated to three OH" ions and also to water,
resulting in a basal spacing of 9.2 A for NO 3 '. Morioka and coworkers39 reported
synthesis o f structurally similar zinc hydroxy acetate from mixing ZnO and Zn(CH 3 C 0 2) 2
in water and allowing the suspension to age for 24 h before filtering, washing, and
drying. This class of layered materials showed interesting anionic exchange reactions

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

with the homologous series of carboxylates with the basal spacing varying linearly with
chain length.39

(b) Hydroxy sub-lattice modification.

(1) Another class of brucite based anionic clays, the layered hydroxy salts, is
made by creation of vacant hydroxyl sites in the octahedral geometry. When a vacant site
exists, the octahedral geometry is then completed by ligation of an anion to balance the
charge within the interlayer space. Nickel based layered hydroxy salts like
Ni3 (OH)4 (NC>3 ) 2 have been synthesized. An illustrative structure is shown in Figure
1.1.3.2. Layered hydroxy salts of Cu2+ like Cu2(0H)3(CH3C02) -FLO containing acetate
anions in the interlayer spaces together with water molecules have been synthesized and
fully characterized.40

o H

o H

O H

O H

o
oH
/ \

O H

O H

O H

Figure 1.1.3.2 A schematic representation of the structure of a layered hydroxy salt


Ni 3 (0 H)4 (N 0 3 )2 . The O of the nitrate anion occupies lattice sites where OH groups are
missing.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

10

Jim'enez-L'opez and coworkers41 performed chemical characterization and X-ray


absorption studies on Cu2 (0 H)3 (CH3 C 0 2 ).H2 0 and its derivatives and they identified two
different Cu-Cu distances indicating the presence of two distinct Cu11crystallographic
sites, probably from distorted octahedra. Yamanaka and coworkers4243 showed that
Cu2(0H)3(CH3C02) H20 can participate in anionic exchange reactions just like HDSs.
Copper hydroxy nitrate, Cu2 ( 0 H)3N 0 3 (naturally occurring mineral gerhardtite) has been
synthesized and characterized with the results suggesting its structure to be consistent
with that of a basic salt.44'45 The structure of gerhardtite shows two distinct coordination
patterns for Cu2+ ions: one copper site is coordinated to four OH and two O (NO 3'), while
the other copper site is coordinated to four OH, one O belonging to NCV, and another
OH. The octahedral pseudocells formed by the metal coordination are deformed with the
four nearest ligands almost in a square planar geometry and the two other ligands at
longer distances. This is attributed to the Jahn-Teller distortation, arising from the uneven
occupation o f the 3dx2-y2 orbital for Cu2+.46

Tanaka and Terada,47 postulated several mechanisms by which these basic layered
hydroxy salts can be formed. According to their postulation one o f the possible
precipitation reaction is given in equation 1.1.3.1 below where copper nitrate reacts with
NaOH as shown;

4Cu(N0 3)2 (aq) + 6NaOH (aq) -> Cu(N03)2-3Cu(0H)2(s)+ 6NaN03 (aq)

1.1.3.1

The product of this reaction Cu(N 0 3 )2 -3 Cu(0 H)2, represent gerhardtite, Cu2(0H)3(N03).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

1.1.4

Motivation for the study


A key property possessed by anionic clays is their ability to reversibly exchange

the interlayer constituent anions. Exchange reactions are general thought to follow a
purely topotactic chemical pathway, conserving the intralayer structure during the course
of the reaction.30 In some cases, however, exchange reactions have been shown to follow
dissolution-reprecipitation mechanism where the host phase dissolves and re-precipitates
containing a different anion from the one initially occupying the gallery space.

The structures, composition, and synthetic preparative methods for layered salts
described here provide ways of generating a wide array of materials with appropriate
utility in exchange reactions. By incorporation o f anions of varying size, shape, and
charge, we anticipate the ability to fine-tune the compositional and structural make-up of
the anionic clays in a quest to come up with new materials with improved electronic,
optical, magnetic, and catalytic properties. Examination of the crystal structure via
powder X-ray diffraction (PXRD) data will be pursued to provide fundamental insights
into these nanostructured materials. In our laboratory, the ultimate long term goal is to be
able to design experimental procedures to perform anionic exchange reactions while
being able to monitor the solid structure evolution in the parent layered material and
anionic release profiles as a function of reaction time and extent.
Systematically probing the following issues will provide insights into the primary
structure and reactivity of these compounds. The primary objectives o f the work
described in this dissertation are:
(1) Characterization ofLHSs and HDSs. We will focus on the fundamental basic
characterization following the synthesis of these compounds. This will afford us a

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

12

detailed understanding of the chemistry related to these compounds. Thus the use of Xray diffraction (XRD), Fourier transform infrared (FTIR), diffuse reflectance
spectroscopy (DRS), thermal gravimetric analysis (TGA), ultraviolet-visible
spectroscopy (UV-VIS), and elemental analysis will provide valuable information on the
structure and chemical reactivity of these materials.

(2) Exploration o f the role that the structure o f the metal hydroxide layer plays in
controlling reactivity. The effect of the identity and relative stoichiometry of metals
present in the HDSs on initial structure and subsequent anion exchange kinetics and
thermal degradation will be the primary focus. In addition, reactivity o f corresponding
layered hydroxy salts, LHSs, (analogues containing a single divalent metal species) will
be investigated in order to provide supporting data to assist in interpretation of HDS data.

(3) Fire retardancy enhancement. Smectite clays and LDHs have been shown to improve
some o f the physical properties of virgin polymers when employed to prepare inorganicorganic polymer composites.16' 1931-34 By controlling the identity of both the cations and
anions, the loading percentages, and the composition of metals present in the structure of
these layered compounds, we will focus on improving the thermal stability o f selected
polymeric materials. Thus, this work will afford us a systematic way to design effective
nanodimensional layered additives that would give the best results in flame retardancy.
This study will provide us with detailed insights into fundamental mechanistic
issues in anion exchange reactions of this class o f 2-D nanostructured materials while at
the same time being able to screen them for industrial and technological applications.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Ultimately we will be able to utilize the gathered scientific information in the synthesis of
tailored materials for specific applications including among others, improving thermal
stability, and fire safety of polymers, and drug delivery systems.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

14

Chapter 2

Hydroxy Double Salt Anion Exchange Kinetics and


Thermodynamics: Effects of Precursor Structure and Anion
Size

2.1

Introduction

2.1.1

Anionic Exchange Kinetics


A variety of layered inorganic/organic hybrid compounds with nanodimensional

interlayer spacings that contain exchangeable anions have been developed for
applications such as catalysts, 1 2 catalyst precursors, 3"5 catalyst support material, 6
adsorbents, antacids, 7 ion exchangers, 8 9 water purification , 48' 51 fire retardancy,16' 19,31'34,
52' 54

and controlled release o f anions for drug delivery . 55"60 For instance, polymer

nanocomposites containing modified layered clays have been shown to provide


enhancements in physical properties of polymers such as increased tensile strength,
tensile modulus, flexural strength, thermal stability, and corrosion protection owing to
their structural morphology . 2 7 ,6 1 ' 67
In particular, layered double hydroxides (LDHs) and basic metal hydroxides have
been the subject of numerous studies. 5 6 ' 5 7 , 68

79

LDHs can be considered complementary

to cationic, or smectite clays and the ability to exchange intercalated ions has been a key
factor in optimizing these materials for different applications. These layered materials
containing 2-D nanometric confinements capable of hosting different anions and
participating in nanoscale reactions, are useful in the synthesis of a variety o f tailor-made
compounds for various applications. Organic and/ or inorganic moieties dispersed in the
ensuing nanometric domains display unique physical and chemical characteristics. The

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

15

identity of the metal ions controls the details o f the intralayer structure. The
exchangeable anion, An'; can either be monovalent, for example, C f, NO 3 ', or CH 3 COO',
or divalent, such as SO4 2 or CO3 2'. Different synthetic routes for obtaining HDSs have
been reported and there have been several studies that characterized anion-exchanged
materials. 1 7 26 37 38 80
While there has been some work reported on the synthesis and characterization of
HDSs, no studies to date have focused on the kinetics and mechanism of the anionic
exchange reactions. However, in the case o f LDHs and other related lamellar materials,
time resolved energy-dispersive X-ray diffraction, (EDXRD), has been utilized to obtain
in situ kinetic data on the evolution of the solid phase structures during intercalation
and/or anion exchange reactions. 81' 87 Anionic exchange reactions are generally viewed as
topotactic solid-state methods of modifying existing layered materials while maintaining
their overall structural integrity. This is possible due to the rapid diffusion of guest
anions into and out of the 2-D interlamellar spaces. Several factors including structural
morphology, intrinsic and extrinsic defects, lattice vacancies, lines and planes of
deformation, interstitials, and grain boundaries play an important role in the reactivity of
these materials. 30
Topotactic exchanges can exhibit different mechanisms, or staging, for the
evolution from starting material to the completely exchanged product, as shown
schematically in Figure 2.1.1.1. First-order staging refers to the process where all layers
begin to exchange simultaneously and 0 0 / features corresponding to both reactant and
product will be observed at intermediate times, leading to a mixture o f the two phases.
Second-order staging refers to intermediate structures with alternating layers of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

exchanged and unexchanged material and is observed less commonly in these materials.
This phenomenon, commonly referred to as interstratified staging, would result in a new
series o f 0 0 / x-ray diffraction peaks (basal reflections) at 2 0 positions corresponding to
the sum (di+d2) o f the interlayer space o f the starting material and the fully exchanged
product.

mwmk


S e c o n d - o r d e r s ta g in g

itoSS3S


i
M ix tu r e o f 2 p h a s e s

Figure 2.1.1.1 First-order staging o f replacement o f initial anion, indicated in this


schematic representation in red, with larger (blue) anion leads to a mixture o f two phases
(right). The PXRD pattern will have two sequences in basal spacing, di and d2. Secondorder staging leads to alternating layers and an extra set o f basal reflections
corresponding to (di + d2).
In analogy with LDH compounds, two key factors in HDS anion exchange
reactivity are expected to be reactant structural parameters (metal ion identity and
intralayer structure) and the identity/structure o f the replacement anion. In the work
reported here, ex situ analyses are used to characterize the exchange reaction o f acetatecontaining HDSs exposed to aqueous solutions containing replacement anions. Proton
nuclear magnetic resonance (*H NMR) spectroscopy is used to follow the kinetics of
acetate release following exposure o f three model compounds to n-alkyl carboxylate
anions and powder x-ray diffraction is utilized to obtain complementary information on

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

17

the structural transformation of the HDS. The results obtained here provide the first
detailed view of kinetics and mechanisms of HDS anion exchange reactions.

2.1.2

Thermodynamics of Anionic Exchange and Chain Packing Modes


Several methods of making layered hydroxide salts have been developed among

others, co-precipitation, 88' 92 anionic exchange93' 95 and rehydration o f metal oxides


obtained following calcination of preformed layered salts taking advantage of the so
called memory effect .7 0 , 96 However, anionic exchange has emerged as the most
convenient way of modifying the interlayer space of these layered inorganic-organic
hybrids while preserving their intralayer structure. Despite the extensive application of
the anion exchange method in modifying the nanometric galleries, very little is known in
relation to the factors governing these topotactic solid-state transformations. Monitoring
the evolution of the exchange products in situ or ex situ by non-invasive methods like
EDXRD, NMR, XRD, and Fourier transform infrared can help elucidate thermodynamic,
kinetic and mechanistic parameters involved during these reactions in which the parent
materials preserve their overall structural integrity.
The rate of anionic exchange is dependent on various factors including the
incumbent anion/host interaction, guest anion/host interaction, solvation energies of both
anionic species, and the activation energies for the formation of respective exchanged
product phases. Varying the reaction conditions greatly affects the morphology of the
exchanged product as interlayer anions strive to adopt energetically favored orientations
within the galleries. Ranges of anionic orientations are possible such as monolayer (fully
intergiditated overlap), bilayer (partial or no overlap with anions tilted relative to the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

metal hydroxide sheets), horizontal orientation, and mixed phases. Anionic exchange in
HDS materials has been extensively studied, 7 , 9 9 7 ' 100 however, only a few compounds
containing a mixture of monolayer and bilayer phases have actually been synthesized and
characterized. This phenomenon is critical in cases where the ratio of the monolayer and
bilayer phases needs to be fine-tuned to attain desired physico-chemical properties such
as controlled magnetism. In layered inorganic/organic hybrids, magnetism is perceived to
originate in the intralayer structure.20 Incorporation o f anions of different shapes and
varying orientations results in subtle changes in the inorganic layer, hence the observed
perturbations in magnetic behaviors. Thus the interlayer occupancy of these hybrids
serves to control and/ or switch magnetic properties.

101

9 0 9 9 109

Fujita and coworkers

have

reported drastic changes in magnetic behavior of layered copper hydroxides following


reversible monolayer and bilayer anionic arrangements in the galleries. The objective of
this study is to examine the effect of experimental parameters on the reaction rate, extent
of anion exchange, anion orientation, and packing density. Parameters to be varied
include: the ratio o f the guest anion to the precursor, the reaction time, and the reaction
temperature. We hope to gain insights into mechanistic features exhibited by these
anionic exchange processes.

2.2

Experimental

2.2.1

Anionic Exchange Kinetics

Copper acetate monohydrate, (98.0%) [Cu(CH3 C 0

)2 -H2 0 ], zinc acetate,

(99.9%) [Zn(CH3 COO)2], nickel acetate tetrahydrate, (98.0%) [Ni(CH3 C 0 0 ) 2 4H 2 0 ],

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

19

and zinc oxide, (99.9%) [ZnO] were obtained from Aldrich Chemical Co. Sodium salts of
formate, acetate, propionate, butyrate, octanoate, (at least 98.0% pure), HPLC grade
methanol (99.9%), 4-hydroxycinnamic acid (98.0%) [p -tO H ^ l^ C ftC I^ C C h H ]
referred herein as [4Hc], and hexanoic acid (99.5%) were also obtained from Aldrich
Chemical Co. Deutrium oxide [D2 O] (99.9%) was obtained from Cambridge Isotope
Laboratories, Inc. L-glutamic acid monosodium salt was obtained from ICN Biomedicals,
Inc., sodium hydroxide (98.1%) [NaOH], from Fischer Scientific, while FTIR gradepotassium bromide [KBr] was supplied by Alfa Aesar. All chemicals were used without
further purification.
The hydroxy double salts were prepared by following a literature coprecipitation
method39 as follows: zinc copper hydroxy acetate (ZCA) was prepared from mixing
0.41g of ZnO (5 mmol) with 1.00 g of Cu(CH 3 C0

)2 H2 0 (5 mmol) in 10 mL of

deionized water with vigorous stirring at room temperature. The resulting suspension
was allowed to stand for 24 h, after which the precipitate was filtered off and washed
several times with deionized water before drying at room temperature. The zinc nickel
hydroxy acetate (ZNA) HDS and zinc hydroxy acetate (ZHA) were prepared in a similar
way, maintaining the starting molar ratio of ZnO to the corresponding metal acetate as
1:1. In some instances, a second treatment with nickel or zinc acetate was necessary in
order to ensure that there was not any remaining polycrystalline ZnO in the product.
The acetate-containing HDSs or ZHA were mixed with 0.2 M solutions of the
target sodium salt of the exchange anion dissolved in D2 0 . Multiple reaction mixtures
were prepared by mixing 30 mg o f the original HDS powder with 3 mL of the exchange
solution and mechanically shaken for a specified period of time. After a range of time

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

20

intervals, filtering the dispersion halted the reaction. The residue was dried at room
temperature prior to x-ray diffraction analysis and the filtrate was stored in sealed vials to
be used for NMR analysis. Complete exchange of the acetate from ZCA was also
attempted by exposing ZCA to exchange anion solutions for 48 h, filtering, and repeating
the process with a fresh exchange solution.
Powder X-ray diffraction patterns were obtained using a Rikagu diffractometer
operated in para-focusing Bragg-Bretano configuration, with a Vi divergence slit (DS),
V scatter slit (SS), 0.15 mm receiving slit (RS), 0.15 mm monochromator receiving slit
(MRS), and a Cu K a (A=1.54 A) radiation source operated at 1 kW. Data acquisition was
performed using step sizes of 0.036720 s. Samples were mounted on microscope slides
using 10% GE 7031 varnish in ethanol after confirming that the varnish mixture did not
perturb the observed patterns. This was confirmed by running control experiments
without varnish (horizontal mount) and comparing with those glued to the holder (vertical
mount). Peak positions, used to characterize the interlayer d-spacing, were determined by
fitting the data to pseudo-Voigt functions using XFIT103, removing the contribution o f the
Cu Ka ,2 wavelength. The instrument response was obtained using the National Institute of
Standards and Technology (NIST) standard reference sample, Si powder (NBS).
Estimation of the crystallite sizes was performed using the Scherrer equation104,
x = kAJ (B-b) cos 0 = kA/ p cos 0

2.2.1.1

where k is a constant (0.9 for powders), p is the full width at half-maximum height o f the
diffraction peak of the material after correcting for Cu Ka ,2 contribution and instrumental
broadening, B and b are the full widths half-maximum intensity for experimental and
instrumental broadening, respectively, while A is the X-ray wavelength, 1.54 A using a

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

21

Cu K 012 source. The Si 111 peak at 28.4 was used for determination o f the instrumental
broadening factor.
The removal of acetate by exchange with other -alkyl carboxylates was followed
using solution JH NMR spectroscopy on a 300 MHz Varian NMR spectrometer. No
evidence for species other than acetate or the exchange anion was observed. Selected
resonance intensities of non-overlapping acetate and the exchange anion peaks were
integrated for filtrates obtained at different exchange times and data were averaged from
three separate NMR spectra. Integrated intensities of the acetate methyl *H NMR peak at
1.84 ppm were also converted to concentrations, normalizing for run-to-run variations by
use of an internal standard of 0.2M methanol in D 2 O. Octanoate concentrations were
determined using the integrated intensity of the triplet at 2.09 ppm corresponding to the
CH 2 adjacent to the carboxyl group.
Fourier transform infrared (FTIR) spectra of the solid materials were obtained
using the KBr method on a Nicolet Magna-IR 560 spectrometer operated at 1 cm ' 1
resolution in the 400-4000 cm ' 1 region. Peak positions were obtained using a peak-fitting
program, ACD/Labs UVIR Processor 7.0 . 105 Diffuse reflectance spectra (DRS) of the
HDS samples were taken on a Shimadzu UV-2501PC in the reflectance mode at room
temperature using BaSCL as a reference, with the DRS absorbance spectra calculated
using the Kubelka-Munk equation. Huffman Labs, Colorado, carried out elemental
analysis using the Atomic Emission Spectroscopy interfaced with Inductively Coupled
Plasma (AES-ICP) for metals determination. ZCA: Cu3 .6 Zni.4 (0 H)7 .6 (CH 3 C0 2 )2 .4 5 H2 0 ,
[Cu (33.6% calc, 35.6% exp), Zn (13.4% calc, 14.8% exp), C (8.5% calc, 9.4% exp), H
(3.7% calc, 3.2% exp)], ZHA: Zn 5 (0 H) 8 (CH3 C 0 2 )2 -4 H2 0 , [Zn (50.1% calc, 48.8% exp),

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

22

C (7.4% calc, 7.6% exp), H (3.4% calc, 3.25% exp)] and ZNA:
Zn3 .2 Nii.5 (0 H)7 .9 (CH 3 C0 2 )i.5 -1 .7 H2 0 [Zn (38.0% calc, 35.7% exp), Ni (16.0% calc,
14.9% exp), C (6.5% calc, 7.6% exp), H (2.9% calc, 3.25% exp)]. The layered acetate
compounds have a nominal exchange capacity o f approximately 3 meq/gram.

2.2.2

Thermodynamics of Anionic Exchange and Chain Packing Modes


An equivalent of NaOH was added to 4-hydroxycinnamic acid in water in order to

prepare a solution of sodium 4- hydroxycinnamate. Dried ZCA HDS (100 mg ) was


mixed with 5 mL of n M 4-hydroxycinnamate solutions for half the desired exchange
period at set temperatures with frequent shaking. The supernatant was decanted and
replaced by a fresh solution for the rest of the exchange time. The exchanged HDS
product was recovered by filtration, washed with water before drying at room
temperature.
Anion-exchange experiments were performed under varying conditions. The
influence of the [4Hc] : [ZCA] ratios (0.025, 0.05, and 0.2 M : 20 g L '1) on anionic
exchange was investigated at 25 and 40 C for exchange times of 24 and 48 h. The effect
of temperature was also examined using a [4Hc] : [ZCA] exchange ratio of 0.05 M : 20 g
L 1 at 25 and 40 C over 48 h. Exchange time effects were also studied for the [4Hc] :
[ZCA] exchange ratio of 0.025 M : 20 g L " 1 at 55 C over 12, 24, and 48 h. Powder X-ray
diffraction patterns and Fourier transform infrared spectra of the solid materials were
obtained as described above in section 2 .2 . 1 .
Thermogravimetric analysis (TGA) of the precursor and exchange products were
performed on an SDT 2960 simultaneous DTA-TGA instrument from 50 to 600 C in air

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

23

using a ramp rate of 20 C/min with sample sizes of 20 1 mg. Quantitative


determination of 4-hydroxycinnamate released following exposure o f the fully exchanged
phase to a 5-fold excess of Na 2 CC>3 solution over 4 days was performed on a Shimadzu
UV-2501PC at room temperature. 4-hydroxycinnamate standards were prepared by
adding 1.5 equivalents of NaOH to the acid form in deionized water. A multiple point
calibration curve was used to estimate the amount of the released anion. The chain
lengths of exchange anions were calculated at the Restricted Hatree-Fock (RHF) 6-31G
level in Spartan.

2.3

Results and Discussion

2.3.1

Anionic Exchange Kinetics


The details of FIDS structure are strongly influenced by the reaction conditions

and only limited structural information has been reported 39 for compounds synthesized
via the method used here. In order to determine the relationship between observed
kinetics and HDSs structural parameters, the acetate-containing layered reactants were
first characterized via PXRD, DRS, and FTIR spectroscopy. HDSs and basic metal
hydroxy intercalates typically form one of two possible intralayer structures, depending
on whether or not copper is present in the material. Figure 2.3.1.1 shows the PXRD data
for ZNA (bottom) and ZHA (top), along with a model for the expected structure.
The 00/ (/ = 1 to 3) basal reflections found in the lower angle region are narrow
and equidistantly spaced leaving no doubt that these materials are layered and posses
high range ordering in the c direction. The peaks at higher angles, which would normally
give information about the intralamellar structure of the layered materials, are

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

24

16000

^12000
-i

8 8000

>>
'to
c
0

c 4000

10
2

20
30
theta (degrees)

40

50

Figure 2.3.1.1 PXRD data for ZNA HDS (lower trace) and ZHA (upper trace);
representation of structure is inset. Insert represents model structure of above compounds
with octahedrally coordinated Me2+ in grey, tetrahedrally coordinated M2+ in black
(above and below Me2+ vacancies), with acetate anions bound to the M2+.

broad and asymmetrically shaped, a result o f turbostratic effects emanating from


translational disorder of the metal hydroxide sheets along the a and b axis, which
destroys the line shapes of the (hkO) reflections. 106 The PXRD data are consistent with
the commonly observed modified-brucite-like HDS structure where the metal hydroxide
layer is comprised of octahedrally coordinated (by OH) metal ions. Up to 25% of the
octahedral sites are vacant, with tetrahedrally coordinated metal ions situated directly
above and below the vacancies. In this model structure the anions coordinate to the
tetrahedral sites, as shown in the inset to Figure 2.3.1.1. The interlayer c/-spacings for
these materials are similar, 13.0 and 13.6 A, comparable to the literature values of
and 13.2 A for ZNA and ZHA, respectively. 39

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

1 2 .8

25

14000

_ 12000
ra, 10000
|
o

8000
6000

'( / )

_c

4000

2000
0

10

20
2

30

40

50

theta (degrees)

Figure 2.3.1.2 PXRD data for ZCA HDS, lines on the bottom indicate positions predicted
for monoclinic P2i/m structure (see text for details); Miller hkl indices for selected
reflections are given. Structure o f isomorphic layered copper hydroxy acetate, calculated
using data from reference 40 is also shown with copper atoms shown in blue, oxygen in
red, and carbon in green (hydrogen positions not shown).

In contrast to the Zn/Ni HDS (ZNA) and Zn hydroxy acetate (ZHA), the coppercontaining HDS is expected to have a different lamellar structure. This is evident in the
PXRD pattern at higher angles shown in Figure 2.3.1.2. Copper-containing HDSs have
been shown to have structures similar to botallackite, Cu 2 (OH)3 Cl, where the copper ions
occupy two crystallographically distinct sites. The reported analysis 4 0 o f the diffraction
pattern of copper hydroxy acetate provides a good comparison for our HDS. In the
copper hydroxy acetate structure , 4 0 also shown in Figure 2.3.1.2, three distinct copper
sites were identified (two in a mirror plane) and the copper ions were found in JahnTeller distorted octahedral coordination with oxygen from the OH or acetate groups. The
acetate group positions were found to be disordered in a 50:50 fashion via a mirror plane

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

26

bisecting the CH3 -C=0 angle, shown schematically in the Figure 2.3.1.2 structure,
calculated from the reported lattice constants. Intercalated water molecules were also
observed between the layers.
While our PXRD data are not sufficient for a complete structural analysis, we
note that it is also consistent with a monoclinic P2]/m structure with nearly identical
lattice constants. The lines at the bottom o f Figure 2.3.1.2 show the expected positions
using CELREF unit cell assignment software, 107 fitting a monoclinic P2i/m to obtain a =
5.60 A, b = 6.13 A, c = 9.38 A, and P = 91.16; indexing for selected peaks is also
shown. The lattice constants for copper hydroxy acetate have been reported as a = 5.60
A, b =

6 .1 1

A, c = 18.75 A, and p = 91.01 . 40 Analysis of the positions of the

0 0

/ peaks

gives a value of 9.3 A for the ^/-spacing in ZCA, comparable to the literature value of 9.4
A39 and significantly smaller than the spacing observed with ZNA or ZHA.

Other structural parameters that may influence the anion exchange kinetics are the
identity of the metal that coordinates with the anions and the structure of how the
carboxylate group interacts with the metal, i.e., the binding mode. ZNA is typically
described as having the nickel occupying the octahedral sites with acetate then bound to
zinc found in the tetrahedral sites .80 However, HDSs substantially enriched in zinc have
also been reported9 with Zn/Ni ratios of 1.8 , similar to the value of 2.1 found here. Such
Zn-enriched structures would not be consistent with zinc occupying only tetrahedral sites.
Diffuse reflectance spectroscopy can be utilized to assess the nickel coordination.
The inset of Figure 2.3.1.3 shows the ZNA spectrum, with the positions indicated for the
four bands previously assigned for octahedral nickel. 108 Absorption bands at 13,708 cm' 1

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

27

ZCA

ZNA

0.12

"ET
0.08

3 0

IZCB

0.04

2-

(/)
.Q
<

250

ZCH

350

450

550

650

750

w a v e n u m b e r (n m )

ZCG

250

350

450
550
650
W avelength (nm)

750

Figure 2.3.1.3 Diffuse reflectance spectra of ZCA, zinc copper butyrate (ZCB), zinc
copper hydroxycinnamate (ZCH), and zinc copper glutamate (ZCG) HDS. Inset shows
DRS o f ZNA HDS. Positions of bands assigned in reference 108 are marked A (V3 :
3 Tlg(P)^- 3 A 2 g(F)),B ('T lg(G)<3 A2 g(F)), C ( 1Eg( G ) ^ 3 A2 g(F)), and D (v2:
3 Ti g(F )< 3 A 2 g(F)).

and 26,041 cm' 1 are consistent with the bands assigned to 3 Tig(F)<3 A 2 g(F) and
3T ig(P ) <3 A2 g(F) transitions, respectively. The Racah parameter calculated using Dou's
equations109 for Ni2 +(Oh) ions in ZNA is 1,024 cm' 1 and it is similar to the value obtained
for the layered hydroxy acetate salt o f Ni2+ (LHS-Ni) (1,041cm'1)108 and P-NiOH (820925cm'1)110 for which the Ni2+ ion is known to occupy octahedral sites. Based on the
composition of our material and the similarity of the DRS spectrum with that of reference
37, we conclude that our material has the typical ZNA acetate structure with acetate
primarily bound to the zinc.
DRS data were also obtained to assist in determining the primary acetate-binding
site in ZCA. Spectra for ZCA, a Zn/Cu HDS where the acetate had been replaced by
butyrate (ZCB) and where the anion had been replaced by glutamate

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

C0 2 CCH2 CH2 CH(NH2 )C 0 2 H), ZCG are shown in Figure 2.3.1.3. While ZCA, ZCH, and
ZCB exhibit similar spectra, there is a significant blue shift in the d-d transitions
observed for the glutamate analogue. Other anions with different functional groups also
showed significant spectral shifts, indicating that the anion is interacting strongly with
Cu2+. While the possibility of acetate groups also binding to Zn2+ sites cannot be
eliminated, the stoichiometry of the ZCA HDS also indicates that acetate must be bound,
at least in part, to the copper sites.

1
; 0.6

^ '8
o0.2
</)

o 0.6

ZCA

o
c

cT
4000

CD

3000
2000
1000
W avenum ber ( c m ')
.

f0.4

ZA

<

ZNA

w
jQ

0.2

ZHA

4000

3500

3000
2500
2000
W avenumber (cm'1)

1500

1000

Figure 2.3.1.4 FTIR spectra of layered compounds and reference metal acetate
precursors.

Having established the primary acetate binding sites, the next step is to determine
the nature of acetate complexation using the infrared carboxylate symmetric, vs, and
asymmetric, vas, stretching modes. According to Nakamotos 111 postulate, the difference
between vs and vas, Avas.s, provides valuable information about the interaction between
the carboxylate anions and the hydroxylated layers. Empirical rules proposed by

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

29

Nakamoto suggested the following order exists in terms of modes o f interaction between
the positively charged ions and the anions: Avas-S(unidentate)> Avas-s (bridging) > Avas_s
(bidentate). FTIR data are shown in Figure 2.3.1.4 and are summarized in Table 2.3.1.1.
Table 2.3.1.1 Carboxylate vibrational frequencies in the precursor metal acetates and of
the layered hydroxides.

Species

V a s

Zn acetate
Cu acetate
ZHA
ZCA
ZNA

1548
1605
1549
1563
1577

(cm'1)

v s

(cm'1)

1457
1420
1394
1410
1408

v a s _s

91
185
155
153
169

(cm'1)

(cm"1) Lit.
Ref.
Literature
A

v as. s

112

185
172

113
108

166

37

In ZHA, where the tetrahedral sites are occupied by the Zn2+ ions, a Avas.s value
of 155 cm' 1 was obtained as compared to 91 cm' 1 for the zinc acetate salt for which a
bidentate bonding mode has been ascribed. We therefore assign the interaction of the
acetate anions in the ZHA system to be unidentate. A Avas.s value o f 155 cm ' 1 compares
very well with 157 cm' 1 determined for Zn(0 2 CCFl3 )2 (SC(NH2 )2 )2 ? where the ligands are
known to be unidentately bound to the Zn2+ions . 112
In ZNA, acetate anions also interact primarily with Zn. A Avas-s value o f 169 cm "1
was obtained which compares very well with 166 cm ' 1 as reported by Rojas and
coworkers. 37 A Avas-s value of such a magnitude also suggests a unidentate bonding
fashion between the acetate anion and the tetrahedrally positioned Zn2+ ions. Avas.s was
found to be 185 cm' 1 for Cu(CH 3 C0

)2 -H 2 0 while a lower value o f 153 cm ' 1 was

obtained for the ZCA HDS. The acetate is known to interact with the Cu2+ ion in a
bridging fashion in Cu2 (CH 3 C 0

)4 -2 H 2 0 ,

11T
which would mean a shift towards bidentate

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

30

bonding in the ZCA HDS. From the proposed binding modes discussed for the various
systems above, the order of relative rates o f intercalation into these isostructurally similar
materials can be predicted to be, ZNA~ZHA>ZCA if binding is the only contributing
factor.

40

<
05

c
o
(C
Q.
CO
"to
U)
03
00

10

15

20

number of carbon atoms


Figure 2.3.1.5 Interlayer spacings for Zn/Cu HDS w-alkyl homologous series where n is
the total number of carbon atoms contained in a given carboxylate. All basal spacings
calculated using an average of 001, 002, and 003 powder x-ray diffraction peaks.
Propionate and butyrate exhibited two sets of basal spacings.

The ability to exchange anions in these materials was first tested by utilizing
typical complete exchange conditions where ZCA was exposed to multiple aliquots of
solutions containing a different n-alkyl carboxylate. Figure 2.3.1.5 shows the variation of
the basal spacing with alkyl chain length. The basal spacing is dependent on chain length
as well as the orientation of anions, which is dictated by the space filling density of
interlayer anions. From the slope of plot o f ^/-spacing versus number of carbon atoms,
the angle of orientation of the carboxylates anions can be found from the relation sin" 1 (a)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

31

= slope/2.54 with the assumption that the alkyl chains exhibit an all-trans conformation
in the interlayer space . 114 The arrangement o f the alkyl chains in the zinc copper HDS is
consistent with a bilayer arrangement and a tilt angle o f 52.9 with respect to the
hydroxylated sheets of ZCA.
The exchange products o f propionate and butyrate each revealed the presence of
two phases as evident from an extra set of peaks in the PXRD patterns at lower
diffractions angles. In all other cases only one product phase was identified. The
additional phase could be due to the co-existence o f both gauche and trans arrangements
o f the exchange anions in the interlayer spacing or possibly the presence o f two phases
with different degrees o f hydration hence different basal spacings, as has been reported
for longer anions exchanged into copper hydroxy acetate . 21 We also note that Fugita and
Awaga reported that they were unable to exchange butyrate with acetate in (Q i 2 (OH)3 X),
citing decomposition 2 0 as a possible difficulty. Clearly, systematic study o f alkyl chain
length effects in the formation o f the product phases is critical for understanding the
mechanism of the exchange processes.
Reference
CH3 OH
Octanoate
Acetate

PPM

Figure 2.3.1.6 H 1 NMR spectrum of the filtrate obtained after 7200 seconds exchanging
the acetate anions with octanoate anions in ZCA.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

32

The kinetics of acetate release upon exposure to exchange anions was monitored
via 'H NMR analysis of filtrate solutions. Representative data are shown in Figure
2.3.1.6. 'HNM R intensities from the original anion (1.84 ppm) and the exchange anion
(2.09 ppm) were integrated and subsequently converted to a dimensionless quantity, the
extent o f reaction (a). Based on models for exchange kinetics in LDHs, the data were first
tested using the Avrami-Erofe'ev nucleation-growth model, 115 116 of a general functional
form as described below.
a= l-ex p [-(k t)f

2.3.1.1

a ranges from zero at the start of the reaction to unity, the point where the acetate
concentration in solution no longer changes, and is calculated using equation 2 .3.1 .2 :

acetate ( 0 ^acetate

(L)]

2.3.1.2

where I acetate (t) represents the integrated 'if NMR intensity of the CH resonant peak of
the acetate anion at time t and I acetate (t*>) is the integrated intensity after the exchange
reaction is complete. The Sharp-Hancock method 84 was applied to obtain linear plots of
the form
In [-In (1- a)] = m In (t) + m In (k)

2.3.,1.3

from which m and k values can be easily evaluated. The value of m can, in certain cases,
be related to the mechanism of exchange reaction in terms of nuclei growth and or 2 dimensional diffusion, while k is the rate constant for the anionic exchange reaction.
Several other models are available in literature for modeling kinetic data of this
nature, 117 as given in Appendix A, but the 2-dimensional diffusion/nucleation model is
the only one that performed satisfactorily, within the applicable a range of 0.15 to 0.85,
for most cases here. Figure 2.3.1.7 shows the extent o f reaction data for acetate release as

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

33

0 .9 5
0 .7 5
0 .5 5
0 .3 5
0 .1 5
-0 .0 5
0

1000

2000

3000 4000
tim e (s)

5000

6000

Figure 2.3.1.7 Extent of the reaction data for release of acetate ( ) into solution and
corresponding loss of octanoate ( A) from solution obtained for reaction with ZCA HDS.
Acetate and octanoate concentrations were independently obtained from 'H NMR
analysis. Fits obtained with Sharp-Hancock analysis shown as dashed lines.

well as loss of octanoate, CH 3 (CH2 )6 CC>2 ", from solution following reaction with ZCA
HDS at 21.00.5 C. Sharp-Hancock analysis o f the acetate data, Figure 2.3.1.8, gives
k= 0.790.17 x 10"3 s'1 and m = 0.790.03. The value of m gives an indication o f the
extent to which the exchange reaction is diffusion or nucleation controlled. An m value of
about 0.5 would suggest a diffusion controlled reaction where the progression of the
reaction is solely dependent on the rate at which the exchange anion diffuses to the front
o f the growing intercalation phase. The value of 0.8 observed here suggests that both
nucleation and 2-D diffusion dependence contribute to the overall mechanism o f the
exchange reaction.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

34

0.7

0.2
5.5

6.5

7.5

c -0.8

-1.3

1.8
In t

Figure 2.3.1.8 Corresponding Sharp-Hancock plot for release of acetate into solution.

The loss of the octanoate exchange anion from solution is faster than the rate at
which the acetate is diffusing out of the interlayer spacing, evident from the fact that the
a-time curves for the loss of the octanoate and the appearance of the acetate in solution
intersect at an a value less than 0.5. One explanation for this observation would be that,
since the exchange phase grows from the edges of the crystal lattice towards the
morphological epicenter of the host cavities, the displaced acetate anions have to diffuse
against a wave of exchange anions driven by an existing osmotic gradient. However, the
amount of the acetate anions displaced was lower than the amount o f the octanoate anions
intercalated into the host material.
The creation of a larger space in the host material would promote the rapid
diffusion of anions towards positively charged reaction centers resulting in a case where
the kinetically stable product prevails over the thermodynamically stable one for short

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

35

exchange times. Some of the anions in the gallery space may not be bound to the metal
centers but may remain adsorbed on the HDS surfaces after filtration since no washing
was done. To drive these reactions to completion, one would need to add more fresh
exchange anion solution following decantation o f the original reaction supernatant, the
complete exchange process used to obtain the samples analyzed for Figure 2.3.1.5. We
note that elemental analysis of a completely exchanged and washed sample is consistent
with a one-for-one exchange of octanoate for acetate.

0.03

J
o
o
<

0.02

0.02

0.01

0.00
1000

2000

time/s

2000

4000

6000

8000

time /s
Figure 2.3.1.9 Acetate solution phase concentrations as a function of time following
exposure of ZCA HDS to butyrate ( ) and octanoate () anions. Inset shows acetate
release following exposure of ZCA HDS to a formate solution.

The effects of exchange anion on the observed acetate release kinetics are shown
in Figure 2.3.1.9 where acetate concentrations are plotted as a function o f time. The
exchange with butyrate also shows Avrami-Erofe'ev type behavior and the results of
Sharp-Hancock analysis are shown in Table 2.3.1.2. Even though the exchange efficiency

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

was higher in the octanoate anion case, 98%, as compared to 55% for butyrate, the rate
constant for reaching a stable concentration of released acetate was found to be a factor
of two larger for butyrate. Use o f formate as the exchange anion led to a strikingly
different temporal behavior for acetate release as shown in the inset of Figure 2.3.1.9.
These data were not consistent with Avrami-Erofe'ev kinetics and instead followed a
simple zero-order behavior over the measured time range.
Table 2.3.1.2 Kinetic data for exchange reactions.

Exchange anion
Formate
Butyrate
Octanoate
Octanoate
Octanoate

HDS formula
ZCA
ZCA
ZCA
ZHA
ZNA

Temp 0.5(C) M
21.0
N/A
22.0
0.800.05
21.0
0.790.03
22.0
0.620.03
22.0
(0.540.03)b

k (10'3 s1)
0.190.06a
1.70.6
0.80.2
2.60.8
(1.01.73)b

aRate constant determined from zero-order kinetic fitting model. bKinetic parameters
obtained using Sharp-Hancock plot shown in Figure 2.3.1.13 do not adequately represent
full time course of the reaction.

PXRD data were obtained for the solid samples collected during the kinetics
experiments in order to provide insight into the structural transformations taking place
during the exchange process. Selected regions of the PXRD data for reaction of each of
the three exchange anions with ZCA are shown in Figure 2.3.1.10. Reaction times
increase moving from bottom to top, with the time (in seconds) indicated for each PXRD
trace. In each case, the 001 peak from the ZCA HDS can be observed at 20 = 9.5 in the
earliest reaction times (found in the bottom trace of each data set). For exchange with
octanoate, shown on the left side of Figure 2.3.1.10, the only other peaks that are
observed are 00/ (/ = 1-3) reflections from the zinc copper hydroxy octanoate HDS
product. The structural transformation o f the solid material in the formate exchange,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

37

shown on the right side of Figure 2.3.1.10, follows a similar pattern of conversion of
ZCA HDS to a single-phase product.
o cta n oate

butyrate

form ate

360 0
7200
600

540

720

360

360

300

30

20 (degrees)

Figure 2.3.1.10 PXRD data for solid collected at times (s) indicated on the left of the
panel following exposure to each of the three anions. Data are offset for clarity but not
otherwise scaled.

In contrast to octanoate and formate reactions, the exchange with butyrate exhibits
a more complex behavior in terms of solid product phases. PXRD patterns obtained from
butyrate exchange, shown in the middle panel of Figure 2.3.1.10, show two sharp peaks
at 20 values of 6.1 and 6.4, corresponding to basal spacings o f 14.5 A and 13.7 A
respectively, with fluctuating relative intensities. Another broader peak at 20 = 5.4,
corresponding to a Aspacing of 16.5 A, was found to grow with increasing time of
exchange. As noted previously, butyrate exchange led to products with two different dspacings (15.1 and 18.6 A) when carried out under complete exchange conditions in H2 O
at room temperature. The use o f D 2 O for NMR studies may have resulted in the small

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

38

differences in basal spacings if interlamellar water is a factor since hydrogen bonding is


stronger in D2 O than in H2 O . 118
Anionic exchange reactions involve many competitive processes. These may
include the energy needed to disrupt the hydrated structures in solution, electrostatic
interactions between anions and the cations both in solution and in the HDS, and
subsequent energy gained when the exchange anion finally coordinates to the metal
center in the HDS lamellar structure.119 Using the simple Avrami-Erofe'ev kinetic model
to model such multifaceted processes cannot give completely detailed mechanistic insight
but does provide a basis for comparison in cases where such kinetics are followed.
However, both kinetic and thermodynamic factors must be considered.
While butyrate exchange has a larger rate constant, the absolute extent o f reaction
is lower than that observed with octanoate exchange. Greater stability for longer alkyl
chain exchange HDS products has also previously been noted in complete exchange
studies.30 However, there is an energetic cost associated with the larger structural change
o f opening the layers that must accompany exchanges involving larger anions, which
may serve to slow the process. Diffusion rates of the exchange anions may also play a
role in determining the observed kinetics since the value o f m extracted from these data is
consistent with both nucleation and diffusion contributions. However, despite the
formation of more than one product structure in the butyrate reaction, the kinetics still
follows Avrami-Erofe'ev behavior and the larger rate constant (relative to the octanoate
reaction) is consistent with both the diffusion and nucleation aspects o f the reaction.
Additional experiments, including kinetics studies of the reverse reactions and
investigation of additional exchange anions, will be helpful to further elucidate the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

39

factors that control extent of reaction as well as the number of observed product phases.
We note that a preliminary study carried out in our laboratory indicates that when the
butyrate exchange is carried out at 40 C, only one product phase (d-spacing = 15 A),
Figure 2.3.1.11, is formed under complete exchange conditions in H20 . Further work is
in progress to characterize the temperature effects on these reactions.

1600

-*

c
13
o
o
>.

800

CO

c
0
c

400

10

12

14

2 theta (degrees)
Figure 2.3.1.11 PXRD trace for ZCB, solid product obtained after exchanging the acetate
anions with butyrate anions at 40 C. A single-phase exchange product is formed.

The exchange of formate for acetate exhibits significantly different kinetic


behavior compared to other ZCA HDS reactions studied here as well as those reported for
LDHs or other layered compounds. The key difference is likely the fact that formate is a
smaller anion. As the reaction proceeds from the edges toward the center, the interlayer
spacing decreases. This pinching in o f the edges may slow the diffusion o f the acetate
out from between the layers. This uncommon behavior has been observed for certain
solid-state dehydration reactions where desorption at the interface is the rate limiting

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

40

step.118 This result has significance for the design o f these materials for controlled release
applications; the relative size o f the exchanging anions can be exploited to provide
nanostructural changes in the solid that results in either a constant release over some time
period of the target species or, for cases following Avrami-Erofe'ev kinetics, a
deceleratory release.

a
g
o

0.8
0.6

CO
CD

0.5

0.4

x 0.2

+ -

1.5

ID

In (t)
-

0.2
0

1000

2000

3000

4000

time (s)
Figure 2.3.1.12 Extent of reaction as a function of time for acetate anions released from
ZHA exposed to octanoate solution and corresponding Sharp-Hancock plot (inset).

Octanoate anions were intercalated into different host materials under the same
conditions to investigate the effect of host structural morphology on the rate o f anionic
exchange reactions. Figure 2.3.1.12 shows the extent of reaction as a function of time
and corresponding Sharp-Hancock plot for the exchange of octanoate anions into ZHA;
the kinetics data are summarized in Table 2.3.1.2. The acetate release kinetics again
follows Avrami-Erofe'ev kinetics and the reaction proceeds to completion more quickly
than that of the ZCA HDS. This is not surprising given that the initial interlayer spacing

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

41

is larger for ZHA, facilitating diffusion o f the reactant anions, and the binding strength of
the acetate groups, found to be primarily in unidentate coordination, is also weaker.

|o

0.8

L_

CO
CD

0.6

0.5 -

'o
sz 0.4
CD

-0.5 -

X
LU

-1.5 -

-4 *

0.2

ln(t)

1000

2000

3000

4000

time (s)

Figure 2.3.1.13 Extent of reaction as a function of time for acetate anions released from
ZNA HDS exposed to octanoate solution and corresponding Sharp-Hancock plot (inset).
Kinetic parameters derived from Sharp-Hancock analysis do not adequately model the
entire time range. The circles and triangles are for separate reaction batches.
Based on its structural similarity with ZHA, the octanoate exchange reaction with
ZNA HDS is expected to follow a similar behavior. However, the extent of reaction
curve in this case, shown in Figure 2.3.1.13, does not follow the simple deceleratory
behavior of the Avrami-Erofe'ev model. At approximately 500 seconds, the acetate
concentration has reached about half of its final value and appears to be leveling off, only
to begin to increase again more rapidly at around 1000 seconds, suggesting a change in
mechanism for the acetate release. We note that two separately synthesized samples of
ZNA HDS exhibited the initial leveling off o f acetate release at ~ 500 seconds. The
apparent mechanistic change is not as obvious from the Sharp-Hancock plot inset in

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

42

Figure 2.3.1.13 but the fit parameters clearly are not a good representation of the
temporal behavior of the acetate release.

8000
5 6000
c
o

4000

3600

co
c
CD

1800

+-*

c 2000

1440
1080

10

12

2 theta (degrees)
Figure 2.3.1.14 PXRD data obtained for solid materials recovered from the reaction of
ZNA HDS with octanoate. Time of exposure to exchange solutions (in seconds)
indicated for each trace. Plots offset for clarity but not otherwise scaled.

PXRD data are shown in Figure 2.3.1.14 for the solid-state material obtained
during the time where the second stage of acetate release is observed. At 1080 seconds,
two broad features assigned to 001 and 002 o f the product octanoate are observed at 3.8
and 1.1, respectively. It is interesting to note that the original acetate phase 001 peak at
6.8 is no longer evident, even though only about half of the acetate has been released into
solution at this point. In addition, a peak is observed at 20 = 9.0, corresponding to a dspacing of 9.8 A, smaller than the 13.6 A basal spacing of the original HDS. The 9.0
peak disappears by 1440 seconds, suggesting that an intermediate phase may be
influencing the rate of acetate release. At the final time shown in Figure 2.3.1.14, a new

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

43

feature appears at 6 .8 that is not observed at the earlier times. Clearly there is a more
complex set o f structural transformations taking place in the ZNA reaction than would be
expected for a topotactic phase transformation. Investigation of the kinetics with
precursors containing other Zn/Ni compositions will be the subject of future work. We
also note that the solutions remained colorless at all times during the reactions,
suggesting that dissolution/reprecipitation is not a major factor.

2.3.1.1 Conclusions
Anion exchange reactions of HDSs and a related basic zinc hydroxy acetate
compound are shown to exhibit a variety of kinetic behaviors which depend on the
relative size of the anions as well as the structure/composition of the layered precursor.
An unusual zero-order kinetics release was observed for replacement of acetate in ZCA
by formate. The Avrami-Erofe'ev kinetic model can be used to describe rate of acetate
release in the reactions of zinc copper hydroxy acetate HDS with butyrate and octanoate
anions as well as the reaction o f zinc hydroxy acetate with octanoate. In the case of ZCA
HDS, the butyrate reaction reaches equilibrium more quickly than does the octanoate
exchange (kbutyrate =

2 .6

10

s' versus koctanoate =

0 .8

10

' s' ) but the octanoate reaches

a final 98% conversion compared with 55% for butyrate. The butyrate reaction also
produces more than one product structure. The reaction o f ZHA with octanoate proceeds
more quickly than that of ZCA with the same anion, consistent with differences in
structural factors (initial interlayer spacing and predicted strength o f acetate binding).
The reaction of zinc nickel hydroxy acetate with octanoate does not follow Avrami-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

44

Erofe'ev kinetics over the entire reaction period. An intermediate species is postulated to
be the source of this discrepancy based on PXRD data.

2.3.2

Thermodynamics of Anionic Exchange and Chain Packing Modes

=12000-

12000

8000 -

4000

S'
CO

rc

9000

6000

two theta (degrees)

CO

CD

25

30

3000

10

15

20

2 theta (degrees)

Figure 2.3.2.1 PXRD patterns following exchange of acetate by 4-hydroxycinnamate at


the specific conditions; [4Hc] : [ZC A ] concentration ratios/temperature/time; (A ) 0.05 M
: 20 g 1 /7 2 5 C/48 h, (B) 0.05 M : 2 0 g I/V 4 0 C /48 h, (C ) 0.2 M : 20 g I/V 2 5 C/48 h,
and (D) 0.2 M : 20 g I/V 4 0 C /48 h. Empty triangles indicate ZCHb while filled circles
represent the Z C A phase. The insert shows PXRD data for ZC A . Data are offset for
clarity but otherwise not scaled.

The PXRD pattern for ZCA is shown in Figure 2.3.2.1 (insert). ZCA exhibits
strong and sharp

00

/ basal reflections indicative o f a well-crystallized phase revealing the

existence of coherent diffraction planes in the c-dimension. Three monotonic basal


reflections at 20 positions of 9.5, 19.0, and 28.6 were observed corresponding to an
average basal spacing, d, of 9.3 A. Structural details of ZCA have been discussed in

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

45

section 2.3.1. An aromatic carboxylate, 4-hydroxycinnamate, was exchanged into ZCA,


at a [4Hc] : [ZCA] exchange ratio of 0.05 M : 20 g L ' 1 at 25 C for 48 h. The PXRD
pattern of the solid product is shown in Figure 2.3.2.1 trace A and reveals that the
structural integrity of the layered precursor is preserved during the exchange reaction. In
addition to peaks corresponding to the 001 (9.5), 002 (19.0), and 003 (28.6) reflections
of ZCA, the pattern shows a new set of peaks at 4.5, 8.7, and 12.8 which are attributed to
the 001, 002, and 003 reflections of the 4-hydroxycinnamate hydroxy double salt phase.
These basal reflections correspond to an average basal spacing of 2 0 . 2 A.
The chain length of the 4-hydroxycinnamate anion was calculated to be 9.1 A.
The observed basal spacing is consistent with a tilted, non overlapping bilayer anion
arrangement within the galleries as shown in Figure 2.3.2.2a. Several researchers have
reported similar anionic arrangements in the interlayer space o f isomorphic-layered
double hydroxides (LDHs) , 120-122 hydroxy double salts, 123 and layered hydroxy salts. 106
The exchanged solid product can be regarded as a mixture of Z C A and small amounts of
the zinc copper 4-hydroxycinnamate phase (ZCHb) (where the subscript b denotes the
bilayer arrangement).
Using the same [4Hc] : [ZC A ] exchange ratio and reaction time but increasing the
temperature from 25 to 40 C yields a similar exchange product (Figure 2.3.2.1 trace B),
however, the relative proportions of respective phases are reversed. This suggests that
high reaction temperatures greatly enhance the exchange efficiency. The bilayer ZCHb
phase is both kinetically and thermodynamically favored. The 001 reflections of ZCHb at
40 C are less broad than those at 25 C indicative o f an improvement in the crystallinity

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

46

a)

HO

*o2c

H ,0

HO

020

o2c

0 2 C

b)

'02C

o2c

Figure 2.3.2.2 Proposed model structures for the bilayer phase (a) and the interpenetrating
monolayer phase (b) of the exchange products. Hydrogen bonding might exist in the
bilayer arrangement.

for the exchange phase at higher temperatures. A [4Hc] : [ZCA] exchange ratio of 0.2 M :
20 g L' 1 at 25 and 40 C for 48 h was also examined. The PXRD patterns of the exchange
products at 25 and 40 C are similar, and are shown in Figure 2.3.2.1 trace C and D
respectively. In both cases only one phase, the bilayer one of the totally exchanged

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

47

product is observed. Increasing the [ 4 H c ] : [ZC A ] ratio exclusively leads to the formation
of ZCHb at both low and high temperatures. We note here some changes in the intralayer
structure o f the exchanged product, probably a result of mechanical shaking, which might
favor turbostratic disorder.

12000

^
CD

9000

6000

- t '

'c/5
c
CD

-t

3000

10

15

20

25

30

2 theta (degrees)
Figure 2.3.2.3 PXRD patterns following exchange of acetate by 4-hydroxycinnamate at
the specific conditions; [4Hc] : [ZC A ] concentration ratios/temperature/time; (A ) 0.05 M
: 20 g L _1/2 5 C /24 h, (B) 0.05 M : 20 g L V25 C /48 h, (C ) 0.025 M : 20 g 1 /7 4 0 C /24 h,
and (D) 0.025 M : 20 g L 'V 40 C /48 h. Empty triangles indicate the ZCHb phase, filled
circles the Z C A phase while asterisks represent the monolayer, Z C H m phase. Data are
offset for clarity but otherwise not scaled.

Effects o f exchange reaction times were investigated at constant [4Hc] : [ZC A ]


ratios and temperatures. The PXRD patterns o f exchange products following exchange of
4-hydroxycinnamate into Z C A at [4Hc] : [ZC A ] ratios o f 0.025 M : 20 g L ' 1 and 0.05 M :
20 g L "1 at 25 and 40 C over 24 and 48 h are shown in Figure 2.3.2.3. Trace A shows the
XRD pattern for the case where 4-hydroxycinnamate was exchanged at a [4H c]: [ZC A ]
ratio of 0.05 M : 20 g L ' 1 at 25 C over 24 h while trace B shows the result obtained at the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

48

same concentration and temperature but over 48 h. Two phases, Z C A and ZCHb are
observed in both cases. The relative intensity o f the 001 peak of ZCH b to that o f Z C A are
similar for the two cases suggesting that prolonging the exchange time has no significant
effect on the exchange efficiency at this temperature. Reducing the [4Hc] : [ZC A ] ratio to
0.025 M : 20 g L' 1 while maintaining the temperature and reaction times (24 and 48 h)
resulted in samples with PXRD patterns presented in traces C and D o f Figure 2.3.2.3
respectively.
The PXRD patterns of these samples exhibit a series of reflections from ZCA and
ZCHb, and also a new series of broader peaks marked with asterisks in Figure 2.3.2.3.
These new reflections at 20 positions of 6.9 and 13.7 correspond to an average basal
spacing of 12.9 A. This would be consistent with anions adopting a perpendicular, fullyinterdigitated monolayer packing that may have aromatic n-n overlap (Figure
2.3.2.2b) . 124 125 This phase will be referred herein as ZCHm; (subscript m denotes
monolayer arrangement). These results are similar to those observed by Xu and
coworkers following intercalation of elaidate into a Zn-Al layered double hydroxide
(Zn2 Al-LDH-excess elaidate) . 126 However, they observed this phenomenon by using an
excess of the elaidate relative to the anionic exchange capacity (AEC) of the LDH.
Hydrophobic interactions are the driving force for bilayer formation . 126
From the empirical formula of ZCA the AEC was determined to be ~ 300 meq g'1.
A [4Hc] : [ZCA] exchange ratio o f 0.025 M : 20 g L ' 1 (250 meq g'1) means there is not
enough 4-hydroxycinnamate to replace all the acetate anions from the galleries assuming
a one to one exchange. The possibility of 4-hydroxycinnamate intercalated horizontally
relative to the hydroxylated sheets is ruled out on the basis there is no evidence of such

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

49

an anionic arrangement in the galleries. The observed interlayer spacing of 12.9 A is not
consistent with such an arrangement.
Interestingly, the bilayer phase is preferred at shorter reaction times than the
monolayer phase. Insignificant amounts o f the monolayer phase are seen in trace C with
the bilayer being the dominant exchange product. Prolonging the exchange reaction time
to 48 h resulted in an exchange product; trace D, where the monolayer phase is now
dominant. This suggests that there is phase transformation over longer exchange times.
The unavailability of in situ methods for monitoring solid phase exchange products
during the course of the reaction precludes authors from pinpointing with certitude when
the observed transformation occurs. Additional work is necessary.

8000

r 6000
c

o
4000
CO

CD

C 2000

10

15

20

25

30

2 theta (degrees)
Figure 2.3.2.4 PXRD patterns following exchange o f acetate by 4-hydroxycinnamate at
the specific conditions; [4Hc] : [ZC A ] concentration ratios/temperature/time; (A ) 0.025
M : 20 g L_1/40 C/48 h, (B) 0.05 M : 20 g L'V40 C/48 h, (C) 0.2 M : 20 g L'V40 C/48
h. Empty triangles indicate the ZCHb phase and asterisks indicate the monolayer Z C H m
phase while filled circles represent the Z C A phase. Data are offset for clarity but
otherwise not scaled.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

The [4H c]: [ZCA] anionic exchange ratios were varied (0.025, 0.05, 0.2 M : 20 g
L1) at a constant reaction temperature o f 40 C for 48 h. The PXRD patterns for the
exchange products are shown in Figure 2.3.2.4. Three phases, ZCA, ZCHb, and ZCHm are
observed when an exchange ratio o f 0.025 M : 20 g L ' 1 is used (trace A). However,
preferential growth is seen for the ZCHmphase relative to ZCHb at low exchange anion
concentrations. When the ratio is increased to 0.05 M : 20 g L' 1 the bilayer phase, ZCHb,
dominates at the expense of the monolayer phase as shown in the PXRD trace B. Very
little amounts of the precursor material are still present. We note here the exchange anion
concentration is higher than the AEC of ZCA. There is an excess o f the exchange anion
to result with a bilayer arrangement o f anions in the galleries. When the [4Hc] : [ZCA]
ratio is increased to 0.2 M : 20 g L ' 1 while maintaining the reaction temperature and time
at 40 C and 48 h respectively, only the bilayer, ZCHb, phase is obtained as shown in
trace C of Figure 2.3.2.4. Exchange anions in excess of AEC drive the exchange reaction
to completion without formation of ZCHm.
Further experiments were conducted to examine the effect o f reaction time at low
anionic exchange concentrations and high reaction temperatures. Using a [4Hc] : [ZC A ]
exchange ratio o f 0.025 M : 20 g L ' 1 at 55 C reaction times were varied; 12, 24, and 48
h. The PXRD patterns of the exchanged products are shown in Figure 2.3.2.5. In all three
cases triphasic products are observed. For the shorter exchange time o f 12 h, the
conversion of Z C A into ZCHb and Z C H m is very slow. However, interestingly, the
bilayer phase is preferred over Z C H m. Thermodynamically, phases with larger interlayer
spacings are more stable, this being attributed to the fact that there is less electrostatic
repulsion between hydroxylated metal sheets. 30 As the exchange time is increased to 24 h

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

51

a reasonable amount of Z C H m is observed, however, the peaks are relatively broader


when compared to those of ZCHb suggesting less ordering in this phase. After 48 h
conversion of Z C A into ZCHb and Z C H m is not complete, however, the relative amounts
of the exchanged products are higher than observed for shorter times. We note here that
the formation of the triphasic mixture in the exchange process is achieved by use of a low
[4Hc] : [ZC A ] exchange ratio, especially when the value is lower than the A E C o f the

precursor.

5000
d 4000
c

D
O 3000
o

'w 2000
c
0

1000

10

15

20

25

30

2 theta (degrees)
Figure 2.3.2.5 PXRD patterns following exchange of acetate by 4-hydroxycinnamate at
the specific conditions; [4H c] : [ZC A ] concentration ratios/temperature/time; (A ) 0.025
M : 20 g I/V55 C /12 h, (B) 0.025 M : 20 g L_1/55 C /24 h, (C ) 0.025 M : 20 g L_1/55
C/48 h. Empty triangles indicate the ZCHb phase, filled circles the Z C A phase while
asterisks represent the monolayer, Z C H m phase. Data are offset for clarity but otherwise
not scaled.

The calculated interlayer spacing for the exchange products alone cannot
guarantee an unambiguous description of anion arrangements in the galleries. We have

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

hypothesized the intermediate phase, ZCHmto result from a monolayer arrangement of


4-hydroxycinnamate in the galleries. However, from the data collected we cannot
completely rule out the possibilities o f co-intercalation o f acetate anions and the guest
anion, 4-hydroxycinnamate in the same phase. Since the released acetate anions are not
removed from the exchange solution we envision equilibrium to exist between these
anionic species. We expect however, that the equilibrium be shifted towards the
formation o f exchange products. Since acetate anions are smaller they are easily solvated
and it becomes increasingly difficult for them to diffuse back into the galleries once they
are replaced by more hydrophobic and less solvated 4-hydroxycinnamate anions. The
kinetics and anionic exchange mechanisms are controlled by energetics of the reaction
systems. To achieve an exchanged product there is always an activation energy that needs
to be overcome. The host-guest enthalpy is higher for the exchange products since a
larger distance separates the metal hydroxide layers. Also hydrophobic interactions
between 4-hydroxycinnamate anions are greater than acetate-acetate ones hence
stabilization of the interlayer domain.
Fourier transform infrared spectroscopy has been used to demonstrate that the
reaction products contain the guest ions. The FTIR spectra of the ZCA, 4hydroxycinnamic acid, and the exchange products are shown in Figure 2.3.2.6 . The
parent material, ZCA, 4-hydroxycinnamic acid, and the exchanged products exhibit broad
absorption bands at ~ 3400 and 3480 cm ' 1 respectively. These features are attributed to
the metal hydroxyl, water O-H stretch, and the carboxylic acid O-H stretch in 4hydroxycinnamic acid. Sharp bands at 3570 and 3750 cm' 1 in the spectra of the fully
exchanged product and 4-hydroxycinnamic acid may be attributed to the stretching mode

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

53

tt

U>
NV
lO CM

<
W
0M
0 ArA
OS
o

CM

1800

1600

1200

1000

800

600

W avenumber /c m -1
Figure 2.3.2 . 6 FTIR spectra of a) 4-hydroxycinnamic acid, b) fully exchanged bilayer
ZCHb, c) triphasic (Z C A , ZCHb, and Z C H m), d) biphasic (Z C A , ZC H b), and e) Z C A .

Spectra are offset for clarity but not otherwise scaled.

of non-hydrogen bonded O -H groups.40 Weak absorption bands in the 2500 to 2700 cm' 1
may be attributed to the stretching vibrations of the hydrogen bonded O -H group . 127 The
existence of several different functional groups in the guest anion gives rise to FTIR
spectra with many absorption bands for the exchange products and the 4hydroxycinnamic acid. For 4-hydroxycinnamic acid, the bands at 1672 and 1691 cm ' 1
may be due to the u (C -O ) mode of the carboxylic group in either free or hydrogen
bonded environments. Bands in the 1421 to 1627 cm ' 1 region may be attributed to the
aromatic u (C -C ) modes. The 5(CH3) is recorded at ~ 1378 cm ' 1 while n (C -O ) modes
absorb between 1104 and 1327 cm'1. Bands below 1000 cm ' 1 may be attributed to 5(C H )
modes . 127

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Intercalation of 4-hydroxycinnamate into the HDS resulted in FTIR patterns


similar to those o f the carboxylate acid. Ionization of the acid group results in the shift of
the u aSy m ( C = 0 ) and u syrh( C = 0 ) vibrational modes from 1672 and 1691 cm' 1 to 1552 and
1424 cm -1 respectively. Other bands previously observed in the free acid are slightly
shifted following incorporation o f 4-hydroxycinnamate into the layered material. New
peaks at 556, 600, and 731 cm' 1 may be due to the metal-OH translational modes. 1

B ip h a sic (ZCA > ZCHb)


T rip h asic (ZCA, ZCHb, Z C H J
B ip h a sic (ZCHb >ZC A )
Fully E x c h a n g e d ZCHb

90

70

50

30
50

150

25 0

350

45 0

550

Tem perature /C

Figure 2.3.2.7 Thermogravimetric analysis (TGA) curves for the exchanged products.

Thermal decompositions profiles (TGA curves) for ZCA and its exchange
products are shown in Figure 2.3.2.7. The TGA curve of ZCA exhibits multiple weight
losses in the temperature range of 50 - 250 C. ZCA loses both surface adsorbed and
intercalation water in the temperature range of 100 - 150 C as evident from the DTG
pattern showing a pronounced peak in this temperature region. This has been explicitly

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

198

confirmed by FTIR analysis of evolved gases during thermal decomposition of ZCA.

Subsequent degradation steps are assigned to dehydroxyalation of the layered hydroxide


sheets, loss of acetate anions (deanation), and thermo-oxidation of the organic species.
Multiple processes in the temperature regime where loss o f the organic species is
expected may be attributed to several parallel and/ or serial events including
dehydrogenation, decarboxylation, graphitization, ketonization, thermal cracking and
oxidation to CO2 . A total mass loss of 36% was observed and from XRD analysis the
inorganic residues are mainly CuO and ZnO . 128
The thermal decomposition profile of the fully exchanged product (0.02 M : 20 g
L"1 at 25 C for 48 h) is significantly different from that o f the precursor. O f particular
interest is the amount of inorganic residue remaining at the conclusion of the combustion
process. From the determined empirical formula o f ZCA, assuming a one to one
exchange of 4-hydroxycinnamate for the acetate, the expected residual mass is 48%.
However, only 32% residue mass was recovered and this suggests that the amount o f the
guest anions is more than the deintercalated species. The Tagaki group

11

reported a

similar phenomenon, where they observed about 250% intercalation of stearate into a
Mg-Al based layered double hydroxide when compared to its predetermined AEC. They
attributed this to hydrophobic interactions between the exchange anions capable of
resulting in self-assembled aggregates o f the guest anions within the interlayer space.
Further evidence of excess guest anions in the galleries of the HDS is revealed by the
thermal degradation profile of the fully exchanged material showing no loss of water in
the temperature range from 50 - 200 C. Excess hydrophobic anions squeeze out water
molecules from the interlayer space as they form aggregates resulting with very little or

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

56

no intercalated water molecules. To balance the resultant excess negative charge in the
galleries, Na+ ions may be intercalated; however, no attempt was made to determine these
ions in this work.
Another fully exchanged ZCHb phase obtained following exchange at 40 C over
48 h using a [4Hc] : [ZCA] ratio o f 0.2 M : 20 g L' 1 was thermally decomposed and the
collected residue was found to be -14% . Complete exchange of the intercalated 4hydroxycinnamate was attempted by exposing the above sample to a 5-fold excess of
Na2 CC>3 solution for 96 h. Using UV-VIS the concentration of 4-hydroxycinnamate was
found to be 1.46 times higher than would have been if the exchange process was a 1:1.
The TGA profiles of the partially exchanged phases show a somewhat similar behavior to
ZCA in the temperature range o f 50 - 300 C. However, for biphasic and triphasic
compounds residual mass percentages fall between 32 and 64%. This is consistent with
the concomitant existence of acetate and 4-hydroxycinnamate intercalated phases in the
exchange products. If the partially exchanged products contain both anions then their
residual mass percentages are expected to fall between those for ZCA and the fully
exchanged product. Loss of surface adsorbed and intercalated water molecules in the
lower temperature region is observed. However, the amounts released are less when
compared to ZCA further evidence of water molecules being squeezed out of the galleries
with the progression of aqionic exchange. As exchange progresses the interlayer
increasingly become hydrophobic, driving out polar molecules.
We have demonstrated that exchange of 4-hydroxycinnamate into a zinc/copper
containing HDS at different reaction conditions lead to formation o f bilayer and
monolayer phases in varying proportions. Changes in the relative concentrations of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

57

reactants, the reaction temperature, and the reaction time influence the phase distribution
from these anionic exchange reactions. In most cases the bilayer phase is dominant over
the monolayer exchanged phase, contrary to the belief that an anti-parallel fully
interdigitated arrangement would reduce repulsion between anions since the negatively
charged ends are greatly separated while hydrophobic interactions are effectively
retained . 1 2 9 ,130
The reaction pathways govern the formation o f an interpenetrating monolayer,
bilayer or mixed phases. Using a very large excess o f the guest anion exclusively leads to
the formation of the bilayer phase, while moderate concentrations result in partial
exchange. The fully exchanged bilayer phase product is formed at both low and high
temperatures when using excess guest anions an indication that the exchange process is
both thermodynamically and kinetically favored. For guest concentrations less than the
AEC of the parent compound the exchange at low temperatures favors formation of the
bilayer phase for shorter times and monolayer phase for longer exchange times. The
complexity of these reactions prevents us from assigning either thermodynamic or kinetic
effects under these conditions. At high temperatures the activation barrier is easily
overcome from the reactant or product side and the formation o f both the bilayer and
monolayer phase suggest that these species are thermodynamically stable.

2.3.2.1 Conclusions
We have demonstrated that 4-hydroxycinnamate can be incorporated into the
hydroxy double salt, ZCA, to form bilayer and monolayer phases by varying reaction
conditions. Series of XRD peaks for the exchange product phases revealed the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

arrangement of guest anions, 4-hydroxycinnamate in bilayer (ZCHb) and monolayer


(ZCHm) fashion. When the [4Hc]/ZCA ratio is below the AEC, the exchange products
consist of ZCA, ZCHb, and ZCHmphases as inferred from both XRD patterns and FTIR
spectra. Anionic packing in the interlayer space is affected by guest anion concentration,
intercalation temperature, and reaction times. Thermolysis of the fully exchanged product
did not show a mass loss commonly associated with adsorbed and interlayer water in
similar layered compounds. This suggests that the interlayer space is highly hydrophobic
as a result o f high anionic packing density resulting from strong aromatic n-n
interactions.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

59

Chapter 3

Thermal Degradation of Acetate Intercalated Hydroxy Double Salts


(HDSs) and Layered Hydroxy Salts (LHSs).

3.1

Introduction

Layered metal hydroxides with exchangeable interlayer anions can be utilized to


provide nanodimensional structures with tunable physical and chemical properties. 1' 9
These materials are also o f interest as possible fire retardant additives for polymers. Work
in our laboratory, shown in Chapters 4 and 6 , demonstrates that the addition o f copperc-j

r i

containing HDSs or LHSs to poly (methyl methacrylate) or poly(vinyl ester)

lO 'J

I 'll

results in enhanced thermal stability. Significant reductions were observed in total heat
release, as measured via cone calorimetry, and there was an increase in the amount of
char formed compared with the virgin polymer. Reduced copper species (Cu and/or
CU2 O), which may play a catalytic role in the stabilization process, were observed in the
pyrolysis residues. Detailed characterization of the thermal degradation pathways of
HDSs and LHSs in the absence o f polymers is necessary to fully understand the role that
these materials play in protecting polymer composites.
Most literature reports to date have focused on thermal degradation of LDHs to
yield a mixture o f metal oxides. 2 8 , 132-138 Formation of metal oxides for catalysis has been
achieved through thermal decomposition of metal salts or crystalline hydrated organic
salts. 139' 142 Both dynamic and static thermal processes are useful for the preparation of
metal oxides for basic research and technological applications. For instance, ZnO
nanoparticles have been shown to exhibit unique physical and chemical properties with a
wide range of applications as in cosmetics, surface acoustic wave device filters,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

60

photodetectors, and gas sensors. 143 Layered materials with predetermined metal ratios
serve well as precursors for well-dispersed oxides that can be used as homogeneous
multicomponent catalysts. Variations in the conditions employed in solvothermal
processes144' 146 have been used to prepare nanocrystals of varying shapes and sizes.
In this work, a detailed analysis o f the thermal degradation pathways is obtained
for a set o f acetate-containing HDS and LHS model compounds in order to explore the
role of intralayer metal composition.

3.2

Experimental
Zinc hydroxy acetate (ZHA) and zinc copper acetate (ZCA) and nickel zinc

acetate (ZNA) HDSs were prepared and characterized using PXRD and DRS as
previously described. Average crystallite sizes, x, were determined using the DebyeScherrer equation as discussed in Chapter 2. Assignments of known compounds were
made using the Powder Diffraction File (PDF ) . 147 Fourier transform infrared (FTIR)
spectra of the solid materials were obtained as described in Chapter 2. Calcined inorganic
residues were stored in a desiccator prior to making FTIR measurements.
Thermogravimetric analysis (TGA) was performed to determine thermal stability and
degradation pathways using a Cahn TG -131 device in the temperature range of 50 - 600
C at a heating rate of 20 C/min in air, flowing at 85 5 mL/min, with sample sizes of
55.0 5.0 mg contained in quartz cups. The gaseous products of the decomposition
process were analyzed using a Mattson-FTIR interfaced with the Cahn TG-131 device.
Differential thermal analysis (DTA) was performed on a SDT 2960 simultaneous DTATGA instrument from 50 - 600 C using 20 1 mg samples heated between 50 and 600

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

61

C at 20 C/min with air as the purge gas (flow rate, 85 5 mL/min). The identities of the
combustion gases were determined by comparing their FTIR spectra to NIST
standards. 148

3.3

Results and Discussion


ZHA and ZNA are expected to have structures consistent with brucite-like 2 6 38

layered compounds where the second metal occupies tetrahedral sites above and below
vacant octahedral positions. Anions bind to the tetrahedral sites. ZCA has been
hypothesized to have a botallackite-like structure, Cu2 (OH)3 C1, (PDF# 8 - 8 8 ) 147 as
extensively described by Masciocchi and coworkers.4 0 Zinc hydroxy acetate, (ZHA)
zinc copper acetate, (ZCA) and zinc nickel acetate (ZNA) were synthesized as described
above and some of the key structural parameters are summarized below. Additional
discussion can be found in our previous study o f anion exchange kinetics in these model
systems, Chapter 2.
The TGA and corresponding derivatized mass, DTG, curves of ZHA heated in air
from 50-600 C are shown in Figure 3.3.1 A and the differential thermal analysis (DTA)
curve for the corresponding heating ramp is shown in Figure 3.3.IB. The TGA and DTG
curves exhibit two distinct degradation regions: one in the temperature range of 50-150
C and the other in the range o f 150-350 C. A final total mass loss o f 38% is observed.
This compares well with the expected mass loss of 37%, which is based on formation of
ZnO as the final product from the empirical formula o f the ZHA precursor. Combining
the TGA and DTA data with TGA-FTIR experiments provides a detailed view of the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

62

decomposition process. FTIR spectra of the gas phase products from thermal
decomposition of ZHA in air are presented in Figure 3.3.2.

100

2 h-

s?
O''

10

~a
IO

CO
CO
CO

o
. o
DTG
TGA

0
DTA

P-2

50

150

250

350

450

550

Temperature 1 C

Figure 3.3.1 A) TGA and corresponding DTG curves of ZHA from 50 - 600 C at 20
C/min in air. B) DTA curve for ZHA from 50 - 600 C at 20 C/min in air.

Surface adsorbed and inter-crystalline water molecules are lost in the first
degradation stage. A shoulder at about 110 C in the DTA curve, shown in Figure
3.3. IB, is assigned primarily to the loss o f the external surface water molecules while a
stronger endothermic feature centered at 140 C is attributed to the loss of the inter-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

63

gallery hydrogen bonded water molecules. The FTIR spectrum at 121 C confirms the
departure of water is the primary channel during the first degradation step. The expected
weight loss due to simple dehydration, based on having four water molecules per formula
unit, is 11%. However, the weight loss in the first stage is significantly higher,
approximately 18%. This suggests that there may be partial dehydroxylation in this low
temperature regime.

0.03
583
389

0.02

301

II

243

198

121

4000

3500

3000 2500 2000 1500


Wavenumber (cm"1)

1000

500

Figure 3.3.2 FTIR spectra of evolved gaseous products collected at the temperature (C)
indicated on the right of each spectrum during thermal combustion of ZHA in air (20
C/min). FTIR spectra were scaled by dividing by 10 except for those at 121 and 198 C,
which were not scaled. Expected positions of absorption bands for water stretching
modes (a), bending mode (b); CO2 anti-symmetric stretch (c); overlapping acetic acid and
acetone C=0 stretch (d) and CH3 deformation (e), and C-C bending for acetone and C -0
stretch for acetic acid (f) are shown. See Figure 3.3.3 for further detail on acetic acid and
acetone spectra.

The second degradation stage is attributed to the collapse of the intralayer


structure releasing a variety of products. Stable products may include acetic acid
produced via deanation; acetone generated via ketonization o f acetic acid; water which

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

64

may be generated via dehydroxylation of Zn-OH linkages, thermo-oxidation o f organic


species, and ketonization of acetic acid; and carbon dioxide resulting from thermo
oxidation of organic species and ketonization of acetic acid. The evolution of CH3 CO2 H
begins at about 150 C and peaks at around 305 C suggesting that the second stage of
decomposition is mainly due to the loss of the acetate groups. In addition to formation of
acetic acid, acetone, water and C 0 2 are formed at temperatures above 300 C. The FTIR
spectrum o f gaseous products evolved at 328 C along with reference spectra148 of acetic
acid and acetone are shown in Figure 3.3.3; additional spectra obtained between 300 and
400 C are found in Appendix B.

Gas phase products


4----- at 328 C

Acetone

0.6

Acetic acid

JS 0.2

0.2
2000

1800

1600
1400
Wavenumber (cm'1)

1200

1000

Figure 3.3.3 FTIR spectra of evolved gaseous products collected at 328 C during thermal
combustion of ZHA in air, acetic acid, and acetone. Literature 148 acetic acid and acetone
spectra were scaled to fit on the same plot.

Aliphatic carboxylic acids having a-hydrogens undergo bimolecular


decarboxylative coupling (ketonization) yielding ketones.149,150 Two mechanisms for the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

formation of acetone from acetic acid over metal oxide catalysts have been proposed. In
the first mechanism, proposed for oxides of low lattice energy, bulk acetates are formed
first and then subsequently decompose pyrolytically into acetone and CO 2 . 151' 153 This is
promoted by strong basicity (ionicity) of these oxides. In the second mechanism,
proposed for oxides of high lattice energy, the initial step of ketonization involves
adsorptive interaction of acetic acid molecules with the metal oxide surface generating
surface acetate species. Abstraction o f an a-hydrogen atom from the acetate anion
oriented parallel to the surface of the catalyst leads to the formation o f an alkylidene
group which in turn reacts with a neighboring carboxylate and hydrogen to form acetone,
water, and CO2 .149150 In order to evaluate these potential mechanisms, solid residues
were extracted at different points along the heating ramp and analyzed via XRD and
FTIR spectroscopy.
XRD patterns o f the inorganic residue collected after heating ZHA in air at a ramp
rate of 20 C/min up to the indicated temperatures are shown in Figure 3.3.4.
Calcination to 150 C yields a mixture of ZHA and wurtzite ZnO (P 6(3) me; PDF# 361451)147 marked with open diamonds. The observation of ZnO reflections after heating
to this temperature is consistent with the previously noted extent of water loss beyond
that which is expected from the number of water molecules per precursor formula unit.
Some dehydroxylation of ZHA to form ZnO and water may be occurring in this low
temperature regime. However, the existence of 00/ reflections of ZHA to the third order
at temperatures as high as 250 C, implies that some o f the ZHA phase is thermally stable
throughout the first stage of mass loss. Surprisingly, the interlayer spacing increased by

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

66

1.7 A with temperature, probably a result of the rearrangement of the acetate anions in
the gallery space following the loss of inter-crystalline water molecules.

10000
D
-i*

8000

D
O

6000

&

4000

200 0

> -_ a w V _ 2 0 0

10

20

30

40

50

60

70

2 theta (degrees)
Figure 3.3.4 XRD analysis of residual products collected at various times during thermal
combustion of ZHA in air (50 - 600 C). Samples were heated at 20 C/min to the final
temperature shown on the right. ZnO expected peak positions are shown with sticks on
the bottom of the figure; Miller indices are labeled on the top.

Table 3.3.1
Peaks
20 ()

31.9
34.5
36.4
47.7
56.8
63.1

ZHA Thermal Residue: Relative Intensities of ZnO X-Ray Diffraction

hkl

Relative Intensity
ZHA (600 C)a Polycrystalline ZnOb Lit. ZHA heated
At 600 Cc
100
0.55
0.57
0.1
002
0.67
0.44
1.0
101
1.00
1.00
0.3
102
0.27
0.23
0.2
110
0.33
0.32
103
0.36
0.29

This work; sample heated at 20 C/min in air to final temperature indicated.


b PDF# 36-1451, ref. 147.
cEstimated from data presented in ref. 39; sample heated for 60 min.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

67

Table 3.3.2

Average crystallite sizes o f metal oxide phases

Crystallite
Precursor
ZnO
250 C
400 C
300 C
ZHA
322 24 400 36 419 39
ZNA
159 11 145 13 188 17
ZCA
338 68 361 45 293 54

sizes (A)
CuO
600 C
600 C
585 56
482 48
496 62 378 28

NiO
600 C

205 50
-

Calcination of ZHA at higher temperatures leads to formation of polycrystalline


ZnO, peak positions and relative intensities for the sample heated to 600 C are found in
Table 3.3.1. Heating between 300 and 600 C yields wurtzite ZnO as the only
polycrystalline phase. Average ZnO crystallite sizes, as determined from the width o f the
100 reflection, are listed in Table 3.3.2. The crystallite size increases with temperature
from 322 22 A after annealing to 250 C to 585 56 A upon heating to 600 C.
Relative intensities of the ZnO reflections exhibit a slight propensity for growth in the
002-lattice dimension. The 002 peak is more intense than the 100 peak, a reversal that
expected for polycrystalline ZnO.147 Morioka and coworkers reported a more pronounced
preference for crystalline growth along the 002 lattice direction when annealing ZHA at
600 C for 60 min; their data are also summarized in Table 3.3.1.39ZnO microstructure
derived from ZHA is thus sensitive to the thermal degradation protocol employed.
FTIR spectra of the residue obtained after calcinations of the ZHA to specified
temperatures and that of bulk ZnO exposed to acetic acid vapor at room temperature for
2 h, identified as ZnO-Ac, are shown in Figure 3.3.5. The FTIR spectrum obtained after
heating ZHA to 250 C can be assigned in analogy to acetate adsorbates on ZnO.154A
weak but sharp M-O absorption band is observed at 620 cm'1, weak and broad M -0 and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

68

5(CH3) absorption bands are found at 700 and 1030 cm'1, respectively and broadened
symmetric and anti-symmetric C-O stretches occur at around 1420 and 1560 cm'1,
respectively. The shift in position o f the C -0 stretches, in comparison to the ZHA
precursor, is consistent with the increase in interlayer spacing observed in the XRD data
in Figure 3.3.4, ascribed to reorganization o f the anions in the galleries. The most intense
bands, assigned to the C -0 stretching modes, persisted in samples heated to 350 C,
beyond the temperature where evidence for ZHA layered structures were observed in
XRD analysis o f residues. The combined FTIR spectra and XRD data for samples heated
to 300 and 350 C suggest acetate in contact with ZnO.

0.8
Z n O -A c
13

s.

0.6

0.4

400
350

300

.Q

<

0.2

250
ZHA

2000

1700

1400

1100

800

500

W avenum ber (cm '1)

Figure 3.3.5 FTIR spectra of residual products collected after heating ZHA at 20 C per
min in air to the indicated temperature (C). Reference spectra for ZHA precursor
(bottom trace) and for ZnO exposed to acetic acid vapor (top trace) are also provided.
Dotted vertical lines indicate absorption band positions for O-C-O anti-symmetric (a) and
symmetric stretch (b), respectively, deformation mode for (8(CH3)) (c), and Z n-0 stretch
(d) in the precursor ZHA.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

69

A shoulder at 1460 cm' 1 in the FTIR trace at 250 C is also observed in the ZnO-Ac
spectrum and this is consistent with adsorption of acetate groups on the ZnO surface
during thermal decomposition. Peak assignments for the acetate anions bound to the ZnO
surface are as follows; 1554 cm' 1 (anti-symmetric O-C-O stretching), 1460 cm ' 1 (5 (CH3 )),
1428 cm ' 1 (symmetric O-C-O stretching), and 1037 cm ' 1 (8 (CH3 )).
The similarity between FTIR spectra o f ZnO-Ac and bulk zinc acetate precludes a
definitive assignment to either species. However, the absence of polycrystalline zinc
acetate phases in the XRD patterns of the inorganic residues in this temperature regime
suggests that ketonization occurs via surface adsorption of acetic acid over ZnO. The
basic nature of the ZnO surface promotes surface adsorption o f acidic molecules
providing catalytic sites over which ketonization progresses.
Loss of surface adsorbed and inter-crystalline water molecules occur primarily in
the temperature range of 50-150 C together with possible partial dehydroxylation of the
precursor ZHA. In the temperature region o f 150-250 C simultaneous dehydroxylation
and deanation results in the formation of ZnO, ZnO surface adsorbed acetates (ZnO-Ac),
acetic acid, and water. At temperatures higher than 250 C, in addition to products
mentioned above, evolution of CO2 and acetone is observed. This is attributed to the
thermal decomposition of organic species and possible ketonization during the thermal
decomposition process of ZHA in this temperature regime.
TGA and corresponding DTG curves of ZNA in air are shown in Figure 3.3.6A
while the DTA curve is shown in Figure 3.3.6B. ZNA is structurally similar to ZHA, thus
would be expected to show a similar thermal degradation pathway.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

70

100

0.8
n
P
o'V)
V)

CO

TGA

0.2

DTG

DTA

50

150

250

350

450

550

Tem perature 1 C

Figure 3.3.6 A) TGA and corresponding DTG curves of ZNA from 50 - 600 C at 20
C/min in air. B) DTA curve for ZNA from 50 - 600 C at 20 C/min in air.

One difference between these two precursors is the more evident third degradation step
observed in the ZNA TGA data. For ZNA in air, a total mass loss o f ~ 8% (calc. 6%
based on elemental analysis of (precursor) is observed between 50 and 150 C and is
attributed to the loss of physisorbed and intergallery water molecules. This is consistent
with a DTA endothermic feature at around 140 C suggesting absorption of heat

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

71

necessary to evaporate water molecules. A significant mass loss (-10 %) in the 150-250
C region is attributed to the loss of water and acetic acid.
Multiple weight losses totaling 17% are observed in the region of 250 to 350 C
with simultaneous loss of water, acetic acid, and acetone in the earlier stages followed by
the evolution of CO 2 at temperatures above 300 C as shown in the FTIR spectra of TGA
evolved gases shown in Figure 3.3.7. A weak DTA endothermic feature at 230 C and a
vigorous exothermic peak at 340 C in Figure 3.3.6A correspond to the complete thermal
decomposition of ZNA resulting from the loss o f OH ions bound to Ni2+ and the
decomposition o f the acetate anions respectively.36

577

392

w, 0.02
<D
0

304

JD
u_

242

C
CD

1 0.01

<

4000

3500

3000 2500 2000 1500


Wavenumber (cm'1)

1000

500

Figure 3.3.7 FTIR spectra of evolved gaseous products collected at the temperature (C)
indicated on the right of each spectrum during thermal combustion of ZNA in air (20
C/min). FTIR spectra were scaled by dividing by 10 except for those at 120 and 197 C
which were not scaled. Expected positions of absorption bands for water stretching
modes (a), bending mode (b); CO 2 anti-symmetric stretch (c); overlapping acetic acid and
acetone C=0 stretch (d) and C H 3 deformation (e), and C-C bending for acetone and C -0
stretch for acetic acid (f) are shown.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

72

8000

6000

JuwJL lJ L I -

________ A__ A.400

-A

>, 4000

>^300
n.,250

(/)
c
0

200

150
002

0 I ^

003

V i A-

10

'

20

30

40

50

A ^ zna

60

70

2 theta (degrees)

Figure 3.3.8 XRD analysis of the inorganic residue from heating ZNA in air to
temperatures indicated; trace for the residue at 600 C shows ZnO reflections (0) and NiO
reflections ().

Data obtained from XRD analysis o f solid residues obtained at various points in
the heating ramp for ZNA are shown in Figure 3.3.8. At 150 and 200 C the XRD
patterns indicate the presence o f the polycrystalline layered ZNA phase. After heating to
250 C, ZNA and ZnO (marked in empty diamonds) are observed. Complete collapse of
layered structure occurs at 300 C yielding polycrystalline ZnO and NiO (marked in
empty squares; PDF# 47-1049).147 Further heating in the range of 300 to 600 C yields
ZnO and NiO. Average crystallite sizes of the metal oxide phases are summarized in
Table 3.3.2. The crystallite size o f ZnO was found to be 145 13 A after heating to 300
C and increased to 482 48 A after heating to 600 C. The crystallite size of NiO after
heating to 600 C was calculated to be 205 50 A using the non-overlapping 2 0 0
reflection of NiO at 43.2 . The presence o f Ni2+ in the intralayer of ZNA retards growth

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

73

of the ZnO phase, with crystallite sizes observed to be one half or less o f the average
sizes obtained from ZHA at temperatures up to 400 C.
Careful inspection of FTIR spectra o f the gas phase products evolved in the
temperature region of 300 and 400 C, Appendix C reveals some differences in the profile
for evolution o f acetic acid, acetone, CO2 and water from ZNA when compared with
ZHA. Evolution of acetic acid from ZNA begins at temperatures < 250 C. There is a
temperature shift in terms of the relative contributions of deanation versus ketonization
channels for the ZNA relative to ZHA. Examination of the relative intensities of the
carbonyl stretching bands centered at 1788 and 1730 cm' 1 for acetic acid and acetone,
respectively, suggests that for temperatures below 340 C, the production of acetone
relative to acetic acid from ZNA is lower than that observed at the same temperature
from ZHA. Further work is necessary to quantify yields. Combined with the XRD
analysis of solid residues, which exhibit smaller ZnO average crystallite size for ZNA,
the pattern of gas phase product evolution suggests that the presence o f ZnO catalytic
surface is responsible for promoting ketonization.
FTIR spectra of solid residues of ZNA obtained at different points in the heating
ramp and ZnO-Ac are shown in Figure 3.3.9. The FTIR trace for the sample heated to
300 C suggests adsorption of acetic acid moieties on the ZnO surface leading to its
ketonization. CO 2 produced during ketonization may then bind to ZnO forming a metal
carbonate species, as evident from a very broad absorption band (1250 to 1800 cm'1) in
the FTIR trace at 350 C. 150 The continuous release o f CO2 and water at temperatures
above 400 C may be attributed to the thermo-oxidative degradation of organic species
like acetone.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

74

2
ZnO-Ac

:3
03, 1.5
0
O
c
1

400
350

</)
_Q

300

0.5
250

ZNA

0 F=

2000

1700

1400
1100
Wavenumber (cm'1)

800

500

Figure 3.3.9 FTIR spectra o f residual products collected after heating ZNA at 20 C per
min in air to the indicated temperature (C). Reference spectra for ZNA precursor
(bottom trace) and for ZnO exposed to acetic acid vapor (top trace) are also provided.
Dotted vertical lines indicate absorption band positions for O-C-O anti-symmetric (a) and
symmetric stretch (b), respectively, deformation mode for (8 (CH3 )) (c), and Zn-0 stretch
(d) in the precursor ZNA.

TGA and corresponding DTG curves of ZCA in air are shown in Figure 3.3.10A
with corresponding DTA data provided in Figure 3.3.10B. A 7% mass loss in the
temperature range of 50-150 C in air, Figure 3.3.10A, is attributed to the dehydration
process resulting from the loss of surface and interlayer water molecules consistent with a
DTA endothermic feature at about 140 C shown in Figure 3.3.10B. At temperatures
above 400 C, the weight increase is attributed to oxidation of Cu and CU2 O to CuO.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

75

100

TGA

DTG

90
cn

C/3

hlO
13

KD

80

70

60

o
o

12
8

DTA

0
4
50

150

250

350

450

550

Temperature 1 C

Figure 3.3.10 TGA and corresponding DTG curves of ZCA from 50 - 600 C at 20
C/min in air. B) DTA curve for ZCA from 50 - 600 C at 20 C/min in air.

FTIR spectra of gas phase products formed at 121 and 192 C, shown in Figure
3.3.11, reveal that water is the primary product in this temperature range. A second
endothermic degradation step that begins at around 200 C is consistent with both
deanation and dehydroxylation, with water and acetic acid observed in the FTIR spectra
in this temperature range. The DTA curve for ZCA in air shows two overlapping
exothermic peaks between 230 C and 320 C. The final step just above 230 C, is

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

76

0.04

=i0.03
03

Jo.0 2
o
CO

JD

<

0.01

4000

3500

3000 2500 2000 1500


Wavenumber (cm' )

1000

500

Figure 3.3.11 FTIR spectra of evolved gaseous products collected at the temperature (C)
indicated on the right o f each spectrum during thermal combustion o f ZCA in air (20
C/min). FTIR spectra were scaled by dividing by 10 except for those at 121 and 192 C
which were not scaled. Expected positions of absorption bands for water stretching
modes (a), bending mode (b); CO 2 anti-symmetric stretch (c); overlapping acetic acid and
acetone C=0 stretch (d) and CH 3 deformation (e), and C-C bending for acetone and C -0
stretch for acetic acid (f) are shown.

consistent with the loss of organic anions through deanation and/ or thermo-oxidation
leading to the formation of CO2 and H2 O as evident from a sharp weight loss occurring at
240 C accompanied by an exothermic feature in the DTA curve. FTIR spectra shown in
Figure 3.3.11 (with additional spectra provided as Appendix D) show that acetic acid,
water and CO2 are the primary products in the temperature range of 250 to 400 C with
very little formation of acetone.
XRD patterns o f the inorganic residue in air after heating to various temperatures
are shown in Figure 3.3.12. The XRD data for the residue obtained from heating ZCA to
150 C reveal the presence of ZCA as the only polycrystalline product. Heating to 200 C

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

77

yields an amorphous product showing a broad feature at 6 . 6 corresponding to an


unidentified phase. Note that the ZCA layered structure disappears at an earlier
temperature than it does for ZHA or ZNA. Weak, broad ZnO features are observed after
heating to 250 C, along with metallic copper (labeled with asterisks in Figure 3.3.12;
PDF# 4836)147 and Cu20 (open triangles; PDF# 35-1091)147 Simultaneous presence of
metallic copper, Cu2 0 , and ZnO suggest macroscopic heterogeneity. After heating to
600 C ZnO and CuO are the only polycrystalline phases formed.

20000

S'

aL_*J11_*_600

r 15000 c

400

o
10000 -

/w _A -s_300

-2

5000
0

20

40

60

2 theta (degrees)
Figure 3.3.12 XRD analysis of the residue from thermal combustion of ZCA in air; the
temperatures to which ZCA was heated are indicated. Only ZnO (0) and CuO () were
observed after heating to 600 C. Peaks assigned to metallic copper (*) and Cu20 (A)
were observed in samples heated to 250 and 300 C.

Results reported herein are similar to those observed by Guo and coworkers from
the calcination o f Zn[Cu(CN)3 ] . 5 Cu is oxidized to Cu2 0 , which in turn is oxidized to
CuO as a result of direct contact with oxygen. The approximate crystallite sizes of ZnO
and CuO particles obtained from ZCA after heating to 600 C in air are 496 62 and 378
28 A respectively. When fitting the XRD data to extract peak widths for determination

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

78

o f average crystallite size of ZnO, the overlapping ZnO 100 and CuO 110 reflections
were fit simultaneously. However, the overlapping peaks and relatively weak ZnO
reflections lead to larger uncertainty in the extracted crystallite sizes indicated in Table
3.3.2. Within the uncertainty limits, there appears to be little ZnO crystallite growth until
600 C, as was also observed with ZNA, possibly due to phase separation effects.

0.8
400

S 0 .6
CC

350

300

03
_Q
L_
o

3< 0.2

250

ZCA
0

4000

3500

3000 2500 2000 1500


Wavenumber (cm )

1000

500

Figure 3.3.13 FTIR spectra of residual products collected after heating ZCA at 20 C per
min in air to the indicated temperature (C). Reference spectrum for ZCA precursor
(bottom trace) is also provided. Dotted vertical lines indicate absorption band positions
for O-C-O anti-symmetric (a) and symmetric stretch (b), respectively, deformation mode
for (8 (CH3 )) ( c), and Zn-0 stretch (d) in the precursor ZCA.

FTIR spectra of the residues obtained when heating ZCA are shown in Figure
3.3.13. The inorganic residues at temperatures above 250 C are acetate-free. There is no
evidence of either metal-acetate formation or adsorption o f the acetate anions on metal
oxides, consistent with the minimal acetone production observed from this precursor.
Instead, carbonate containing species are seen at temperatures of 300 C and beyond
probably as a result of combination o f CO2 with metal oxides. 150 This might explain the
relative lack of ketonization observed for this system. Metallic copper and CuO have

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

been reported to have no catalytic effect on the ketonization of acetic acid to acetone . 155
Instead, Pestman and coworkers155 observed that CuO promotes formation of
acetaldehyde from acetic acid. In the same work, copper in its partially reduced form,
CU2 O has been shown to promote ketonization. Even though Q 12O is present in the
temperature region over which acetic acid is released, its concomitant presence with
metallic copper and/ or CuO or the existence of bound carbonate species may render it
inactive. In addition, ZCA has the smallest mole fraction of zinc of the three model
compounds investigated, limiting the ZnO surface area available for catalyzing the
ketonization channel.
Surface adsorbed and inter-crystalline water molecules are largely lost in the
temperature range o f 50-150 C. Decomposition o f ZCA via dehydroxylation and
deanation in the 150-400 C temperature region results in the formation o f ZnO, CuO,
and reduced copper containing species, Cu, CU2 O. Acetic acid, water, and CO2 are
evolved during the thermal decomposition at elevated temperatures. Very little acetone is
observed suggesting that calcination o f ZCA does not promote ketonization.
Polycrystalline ZnO and CuO are the observed solid residues remaining after complete
combustion of ZCA in air.
While heating ZCA did not promote ketonization, we note the reduced metal
species observed in the thermal degradation process. The presence of reduced metallic
species has been shown to play an important role in thermal stabilization of polymeric
m

composites.

U 'X

1T1

1^6

1C7

Klenov and coworkers

1*18

reported the appearance of Cu (I)

and metallic copper on the surface of the Cu2+ containing ZnO after catalyst reduction
and they believe these species play an important role in the catalytic synthesis of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

80

methanol from H2 and CO. The formation o f reduced copper is thus important for
enhancing selected chemical pathways.
Some of the key differences in structure between ZHA, ZNA, and ZCA were
discussed in Chapter 2. FTIR spectra suggest that the bonding of the acetate to the metal
centers in the gallery spacing is pseudo-bridging in ZCA with one acetate oxygen bound
to Cu2+ and the other acetate oxygen forming a hydrogen bond with an adjacent OH
group from the metal hydroxide layer, similar to that reported40 for
Cu2 (0 H)3 (CH 3 C0 2 >H 2 0 . ZHA and ZNA spectra were more consistent with unidentate
binding, however, the possibility of hydrogen bonding interaction of the free carbonyl
oxygen cannot be precluded on the basis of the FTIR data. DTA curves from thermal
degradation of ZCA in air reported here exhibit a pronounced endothermic event around
210C, which was not observed for ZHA and barely observed for the ZNA thermal
degradation in air. This is consistent with a stronger anion-cation interaction for the ZCA
HDS as compared to ZHA and ZNA. The departure of the acetate anion without
decomposition requires more energy if it were tightly bound to the metal center as
assumed in ZCA, while less energy will be required for the same processes when the
interaction between the metal center and the anion is weaker, as predicted for ZHA and
ZNA.

3.4

Conclusions
The thermal degradation o f zinc-containing layered hydroxy salt or hydroxy

double salts with interlayer acetate depends upon the structure/composition o f the metal
hydroxide layer. The initial thermal degradation step for ZHA, ZNA, and ZCA is loss of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

intercalated water in the temperature range of 50-150C. Other gas-phase products


observed at higher temperatures include acetic acid, acetone, CO2 , and H 2 O, and the solid
residues contain a mixture of metal oxide phases. Catalytic ketonization of acetic acid to
form acetone is observed for ZHA and ZNA, with slightly higher temperatures required
for the ketonization channel in ZNA. Little acetone was observed in the thermal
decomposition of ZCA. Combined FTIR and XRD analysis o f solid residues extracted at
different points in the heating profile suggests that ketonization occurs via dissociative
adsorption of acetic acid on ZnO surfaces.
ZnO from the thermal degradation o f ZHA revealed a slight preferential growth in
002-lattice dimension. Inclusion of the other metals (Ni and Cu) in the HDS structure
containing Zn eliminated preferential crystal growth in any lattice directions for ZnO and
also served to retard the ZnO crystallite growth. In the case o f ZCA, reduced copper
species, Cu20 and metallic copper, were observed in the thermal degradation of ZCA at
250 - 500 C in an air atmosphere with oxidation to CuO completed by 600 C.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

82

Chapter 4

Thermal Stability of Hydroxy Double Salt (HDS) and Layered


Hydroxy Salt (LHS)/Poly(methyl methacrylate) (PMMA) Composites.

4 .1

Introduction
Polymer nanocomposites have been comprehensively studied for some time and

have generated a lot of interest in the research field of polymer science and material
chemistry. These nanocomposites have been shown to offer improved fire retardancy and
superior physical properties including increased tensile strength, increased tensile
modulus, increased flexural strength, and an appreciable elevation o f the heat distortion
temperature. 1 1 , 6 1 , 159 Previous studies have been focused on layered silicates in form of
11 -l-J

natural smectite clays and layered double hydroxides (LDHs). '

160

Hydroxy double

salts (HDSs) and layered hydroxy salts, (LHSs) a complimentary class of natural clays
containing nanometric domains capable of hosting guest polymer chains in their gallery
spaces are investigated in this work.
HDSs, are represented by a generic formula [(M2 +ix,Me2 +1+x)(OH) 3 (i.y)]+An'
(i+3y)/nzH2 0 , where M and Me are Zn2+ and Cu2+ ions respectively, with An' being the
interlayer charge compensating anion, the acetate, as in zinc copper acetate (ZCA). The
precursor layered hydroxy salt used here, copper hydroxy nitrate, (CHN) has a formula
predicted to be, Cu2 (0 H)3 N 0 3 . These layered materials have appreciable tunable
properties that include variations o f the constituent metal ions in the intralayer domain
and the use o f different interlayer organic and or inorganic fillers. We note a relatively
higher exchange capacity o f the HDSs and LHSs as compared to natural clays a factor
that would promote easy diffusion of polymer chains or monomer into the gallery spaces.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

83

The relatively small basal spacing in ZCA and CHN would limit the amount of
polymer intercalated between the layers of the salt, hence the desire to exchange the
acetate and nitrate anions with longer polymerizable organic anions, methacrylate,
(CH2 =C(CH 3 )CC>2 ) in this study. Methacrylate anions increase the hydrophobicity of the
interlayer space, making the entrance o f monomer or polymer chains easier. This results
in either intercalated and or exfoliated composites, of which the latter is less likely due to
the high flexibility of the metal hydroxide slabs that make the boundaries of these 2-D
nanocavities.
The modification of the ZCA and CHN changes the surface chemistry o f their
hydroxylated metal sheets facilitating the intercalation o f the polymers. A polymerization
reaction between the vinyl end group of the CH2 =C(CH 3 )C 0 2 ' and the monomeric units
is envisioned, leading to a better dispersion o f the polymer in the interlayer spaces and
thus increasing the flame retardancy. It has been reported 10 that the probability o f
obtaining exfoliated nanocomposites is increased by the presence of a polymerizable
double bond in the gallery space o f the clays.

4.2

Experimental
Monomeric methyl methacrylate, [CH2 =C(CH 3 )C 0 2 CH3], copper (II)

methacrylate (97%) [(H2 C=C(CH 3 )C 0 2 )2 Cu], and potassium bromide (FTIR grade)
[KBr] were obtained from Alfar Aesar. The initiator, benzoyl peroxide (BPO), copper
acetate monohydrate (98.0%) [Cu(CH 3 C0 2 ) 2 H 2 0 ], sodium methacrylate (99.0%)
[H2 C=C(CH 3 )C 0 2 Na], and zinc oxide (99.9%) [ZnO] were obtained form Aldrich
Chemical Co. Copper (II) nitrate (98.9%) [Cu(N 0 3 )2 2 %H2 0 ] was obtained from Fisher

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

84

Scientific Co. while ammonium hydroxide

[N H 4 O H ],

was obtained from EM Science,

Merck. All chemicals were used without further purification with the exception o f methyl
methacrylate monomer solution freed from the inhibitor by passing through an inhibitor
remover column acquired from Aldrich Chemical Co.
The layered hydroxy salt, copper hydroxy nitrate (CHN) was made following
already established precipitation methods .47 Copper (II) nitrate [ C ^ N C ^ ^ A ^ O ] , (10.0
g; 43 mmol) was added to 100 mL o f distilled water and the pH o f the resultant solution
subsequently raised to

by dropwise addition of aqueous ammonia. The solution was

allowed to stand at room temperature for 24 h after which the precipitate was filtered off,
washed and dried. The hydroxy double salt, zinc copper acetate (ZCA) was made from
ZnO (0.41 g; 5 mmol) mixed with Cu(CH 3 C0 2 )2 H 20 (1.00 g; 5 mmol) in 10 mL of
distilled water as previously discussed in Chapter 2. The suspension was aged at room
temperature for 24 h filtered, washed and subsequently dried.
Methacrylate anions were partially exchanged for the acetate in ZCA and for the
nitrate in CHN respectively, by mixing the dried precursor powders, ZCA and CHN, with
0.2 M solution of sodium methacrylate. The exchange products were primarily zinc
copper methacrylate (ZCM) and copper hydroxy methacrylate (CHM) with some o f the
precursor phases still present. In a typical exchange reaction at room temperature, 10.0 g
o f ZCA were mixed with 500 mL o f the exchange solution and frequently agitated for 48
h. The solution was then decanted and replaced with a fresh one for another 48 h.
Methyl methacrylate monomer was introduced into a 200 mL beaker together
with the initiator, BPO, (1%) and the modified synthetic anionic clays. A target loading
percentage, x%, was obtained by adding together ( 1 0 0 -(l+x)) g of the monomer,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

g of

85

BPO, and x g of the modified clay. The mixture was initially heated to 90 C with
vigorous stirring until viscous after which the temperature was then lowered to 60 C and
held constant for 24 h. The temperature was then raised to 80 C and held at that
temperature for another 24 h followed by vacuum drying the sample at 100 C for 12 h to
drive off excess monomer. Pure poly (methyl methacrylate) (PMMA) was made in a
similar way except that no anionic clays were added.
Powder x-ray diffraction (PXRD) patterns of the synthesized layered materials
were obtained as described above. Polymer composite samples were pressed into 1 mm
thick platelets, which were then mounted onto vertically oriented sample holders for
XRD analysis. Fourier transform infrared (FTIR) and diffuse reflectance spectra (DRS)
of the anionic clays were determined as described before. Thermogravimetric analysis
(TGA) was done on a Mattson-Cahn TG-131 device in the temperature range of 50-600
C in both air and N 2 using a ramp rate of 20 C/min with samples sizes in the range of
50-60 mg. Differential thermal analysis (DTA) was performed on a SDT 2960
Simultaneous DTA-TGA instrument from 50-600 0 C using a heating rate of 20 0 C/min
in both air and N 2 with sample sizes of 10-15 mg.
PMMA composite samples were compression molded into square plaques of
uniform thickness before cone calorimetry was performed on an Atlas Cone 2 instrument
at an incident flux of 50 kW/m 2 with a cone shaped heater. Bright field transmission
electron microscopy (TEM) images were collected at 60 kV with a Zeiss 10c electron
microscope at Cornell University by David Jiang. The exchange capacity of ZCA was
calculated to be ~ 3.0 meq/gram while an exchange capacity of 4.2 meq/gram would be
calculated for CFIN, from its nominal formula, Cu2 (0 F1) 3 N 0 3 .

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

86

4.3

Results and Discussion


The layered hydroxy salt, copper hydroxy nitrate (CHN), isomorphic to the

mineral gerhardtite, was found to exhibit a lamellar structure as evident from powder xray diffraction patterns, Figure 4.3.1, showing sharp, intense, and equidistantly spaced
00/ (/=0 to 3) reflections. The structure o f gerhardtite and its synthetic analogous have
been extensively studied, 4 4 45 and found to possess a bmcite Mg(OH)2 -like structure,
where Cu atoms form dual wave-like layers at about equal displacements above and
below the median plane. These Cu atoms form distorted hexagonal pseudocells,
exhibiting two different kinds of lattice occupation. At one copper site, the atoms are
coordinated in a square planar fashion to 4 OH ions with two O (of NO 3 ions) completing
the coordination octahedron; while at the other site, Cu forms a similar square planar
structure capped by one OH ion and one O (of NO 3 ).
The basal spacing was found to be 6.9 A, consistent with N O 3 ' anions occupying
the interlayer space with their C 3 rotational axis perpendicular to the c-axis o f the unit
cells, Figure 4.3.1 (insert). Fourier transform infrared analysis, Figure 4.3.2, revealed the
characteristic absorption peaks at 1050 and 1384 cm ' 1 corresponding to the symmetric
and asymmetric stretching of the N O 3 ' ions respectively. The triplet signature centered at
1384 cm' 1 is consistent with free N O 3 ' groups existing in the gallery spaces. When the
nitrate groups are bound to the intralayer their symmetry is reduced from high D 3 h
symmetry to a C2 Vone resulting in the disappearance of the triplet signature. 111 A broad
peak centered about 3450 cm' 1 is a signature of the presence of lattice water, while sharp
peaks at 3550 and 3730 cm' 1 are attributed to the OH stretching modes.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

87

6000

0 i---------------1---------------1--------------- 1--------------- 1---------------

10

20

30

40

50

2 theta (degrees)
Figure 4.3.1 PXRD pattern for (a) Cu 2 ( 0 H) 3 N 0 3 , (gerhardtite) and (b) partially
exchanged Cu 2 ( 0 H) 3N 0 3 with methacrylate anions to form CHM (). Insert shows the
model structure o f gerhardtite, Cu atoms (green) coordinated to OH groups (red), NO 3
anions occupying the gallery space with the C 3 axis perpendicular to the c-direction o f the
lattice cell. Patterns are offset for clarity but otherwise not scaled.

0.8

0.4
o

V)
-Q

< 0.2

CHN

CHM

4000

3500

3000

2500

2000

1500

1000

W avenum ber (cm-1)


Figure 4.3.2 FTIR spectral patterns o f CHN and the exchange product CHM. Spectra
offset for clarity but otherwise not scaled.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

The nitrate anions were easily replaced by methacrylate anions to yield a new
partially exchanged layered salt, primarily consisting o f the copper hydroxy methacrylate
(CHM) phase. The basal spacing increased from 6.9 A to 13.4 A with the stacking
periodicity preserved in the exchange product, indicating that little or no local structural
change occurred in the inorganic layers. FTIR analysis of the exchange product, CHM,
Figure 4.3.2, revealed the presence o f the organic moieties, methacrylate anions, in the
galleries, as confirmed by several bands between 2800 and 3000 cm'1, which are assigned
to the C-H stretching vibrations. A peak at 1260 cm'1 is attributed to the C -0 bending
mode while peaks at 1421 and 1558 cm'1 are due to the C=0 symmetric and anti
symmetric stretching modes respectively. This suggests that the methacrylate anions
occupy the gallery space and are connected to the Cu2+ ions through their carboxylate
ends. The triplet of the NCV is shifted and greatly depleted, an indication that most of the
N O 3'

ions were replaced during the exchange process. The same conclusion can be drawn

from the PXRD of the exchange product, CHM as shown in Figure 4.3.1.
Zinc copper acetate (ZCA) has been studied in our laboratory, (see Chapter 2)
and predicted to possess a structure similar to botallackite, Cu2 (OH)3 C 1, where the
inorganic layers are constituted by both Zn

2+

and Cu

2+

ions with the later occupying two

distinct crystallographical sites to which the interlayer anions are coordinated. On


exchanging the acetate with the methacrylate, PXRD patterns in Figure 4.3.3 show two
distinct sets of basal reflections consistent with the formation of a and p phases as
reported for butyrate in Cu2 (OH)3 (alkanecarboxylate) by Fugita and coworkers.20 The
basal spacings were found to be 12.0 A and 13.7 A for the a and p phases, respectively.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

89

8000
001

5co 6000
c

8 4000
ZCM

C 2000

002

003
ZCA

10

20
30
2 theta (degrees)

40

50

Figure 4.3.3 PXRD data for ZCA HDS (lower trace), and ZCM (upper trace) scaled by a
factor of 2; a ( A) and p () series are shown for the exchange product ZCM.

1200

ri

PMMA-ZCM2%

900

600

&

PMMA-CHM2%

300

10

20

30

40

50

2 theta (degrees)
Figure 4.3.4 XRD patterns of the PMMA composites, PMMA-ZCM2% (upper trace) and
PMMA-CHM2% (lower trace):

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

90

X-ray diffraction (XRD) was used to characterize the layered structure of poly
(methyl methacrylate) (PMMA)-anionic clay composites. Figure 4.3.4 shows the XRD
pattern for PMMA-CHM2%, a composite made by loading the polymer matrix with 2%
CHM modified clay. A relatively weak peak at 20 = 7.5 corresponding to a d-spacing of
11.8 A shows a reduction in the gallery spacing o f the composite as compared to the
layered hydroxy salt. The intensity o f this peak is greatly diminished as compared to the
001 of CHM, an indication of the loss o f structural ordering.
Similar results were observed with PMMA-ZCM2%, a composite made by
loading the polymer matrix with 2% o f ZCM, for which the XRD pattern is also
presented in Figure 4.3.4. A reduced basal spacing of 11.7 A calculated from a rather
weak 00/ reflection peak at 7.6 0 is reported. As already stated above, the reduction in the
basal spacing may have nothing to do with the polymer molecule insertion into the
lamellar framework o f the layered inorganic/organic modifiers. The contraction in the
interlayer spacing could however be attributed to the loss of the interlayer water since the
polymer composites are vacuum dried at 100 C.
The presence of diffuse, less intense 00/ reflection peaks with a broad background
suggests the presence o f some amount o f the intercalated material within the polymer
matrix. Thus, hydroxylated metal sheets provide protection to the PMMA chains, which
may be taken as evidence to suggest polymer chain insertion in the interlamellar space.
However, it should be noted that XRD patterns alone could not give conclusive
morphological structural description o f these composites. Transmission electron
microscopy (TEM) provides direct complementary evidence of anionic clay dispersion in
the polymer matrix. Low and high-resolution images provide information on the quality

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

91

of the dispersion of the modified anionic clays in PMMA-composites. Low magnification


images allow one to determine whether a microcomposite and or nanocomposite has been
formed, while high magnification images would help one to determine whether
intercalation or exfoliation has occurred. TEM images of PMMA-composites PMMACHM2% and PMMA-ZCM2% are shown in Figures 4.3.5 and 4.3.6 respectively.
The low magnification TEM images of PMMA-CHM2% showed the formation of
a microcomposite with the modified anionic clay clumped together suggesting a poor
dispersion in the polymer matrix. On the other hand, high-resolution image of PMMACHM2% neither show delamination nor intercalation. These results are also confirmed by
XRD analysis, Figure 4.3.4, of this polymer composite where the d-spacing is even
reduced after polymerization. However, low magnification image o f the PMMA-ZCM2%
shows better dispersion of the anionic clay in the polymer matrix regardless of

Figure 4.3.5 TEM image at low (100 nm) (left) and high (50 nm) (right) magnification
for PMMA-CHM2% anionic clay composites prepared by bulk polymerization.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

92

Figure 4.3.6 TEM image at low (100 nm) (left) and high (50nm) (right) magnification for
PMMA-ZCM2% anionic clay composites prepared by bulk polymerization.
the presence o f tactoids. This system, if further modified, could result in the formation of
nanocomposites. No intercalation or delamination features are observed in the highresolution image of PMMA composites studied here.

100

CO
CO

co
2

40 -

20

50

150

250
350
450
Temperature PC

550

Figure 4.3.7 TGA curves for A) ZCM, B) pure PMMA, and C) PMMA-ZCM2%, D)
PMMA-ZCM4% composites.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

93

100
80
vO

60

CO

40

20
0

50

150

250

350

450

550

Temperature /C
Figure 4.3.8 TGA curves for A) CHM, B) pure PMMA, C) PMMA-CHM2%, D)
PMMA-CHM4% composites.

Thermogravimetric analysis (TGA) curves o f ZCM (A), pure PMMA (B),


PMMA-ZCM2 (C), and PMMA-ZCM4 (D) samples in air are shown in Figure 4.3.7. The
pure PMMA sample decomposes in three steps in the temperature range of 200-450 C
leaving no residue after being heating to 600 C. PMMA-ZCM2% and PMMA-ZCM4%
begin to lose weight at about the same temperature as pure PMMA. However, the
degradation onset as depicted by, Tio, the temperature at which 10% of the original mass
is lost, is shifted to higher values for the PMMA-hydroxy double salt composites. TGA
curves of CHM (A), PMMA (B), PMMA-CHM2% (C), and PMMA-CHM4% (D) are
shown in Figure 4.3.8 and they show similar results. No significant weight losses were
observed at 100 C, a clear indication of very little or no physically adsorbed water
molecules in these polymer composites. These results are summarized in Table 4.3.1.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

94

Table 4.3.1

TGA data for pure PMMA and its composites in air and N 2 .

Sample

T 10%
a t 50 o/o Char (%)
T50%
Air n 2 Air n 2 Air n 2 Air n 2
0
0
0
PMMA
249 248 333 338 0
PMMA-CHM2% 282 285 386 388 53 50 2
3
PMMA-CHM4% 277 273 394 393 61 55 6
7
PMMA-ZCM2% 289 288 375 373 42 35 1
0
PMMA-ZCM4% 296 287 384 384 51 46 4
3

The degradation of the PMMA composite materials occurs in two steps, with the
first one probably a result of the weak link scission of MMA monomeric units involved
in head to head linkage, loss of the additive organic content, and/ or weak links due to
disproportionation termination producing vinyledene chain ends. Unreacted initiators of
polymerization, in-chain weak links, and free radicals generated from oxidation of the
monomer methyl methacrylate could be initiators for the thermal degradation course of
PMMA and its composites. 161 This accounts for about 30% of the overall weight loss,
which suggests that the first degradation step cannot be attributed solely to the unreacted
monomer units. Also, since the composites were made in an air atmosphere, weak
peroxides and or hydroperoxides links are expected and these lead to a high degree of
chain scission initiation at low temperatures. 162
Chen and coworkers163 reported that a PMMA nanocomposite sample with 30
wt% MgAl LDH (containing dodecyl sulfate) showed a 45 C increment in the T50 value.
We note here that our values are much better than the reported163 even when working at
such low loadings as 2 and 4%. Zinc polymethacrylate gave a similar increment in the
thermal stability at 50% decomposition an indication that the presence of the metal atoms
in these composites plays a significant role in preventing depolymerization. 164

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

95

Chandrasiri and coworkers165 reported a stabilizing effect of SnCU and tetraphenyltin on


the thermal stabilization of PMMA. They suggested that monomer or polymer radicals
combine with Sn-based radicals forming cross-linked char, which then acts as an energy
or mass transfer barrier. The metal ions contained in our PMMA composites could be
reacting in the same way as the Sn-based radical hence the observed thermal stabilization
effect.
The temperature at which 50% of the polymer composite is lost, T5 0 , is 42 and 51
C higher for PMMA-ZCM2% and PMMA-ZCM4% than for the pure PMMA
respectively. AT50 , the change in T5o of the pure PMMA and composite samples is 53 and
61 C higher for PMMA-CHM2% and PMMA-CHM4% when compared to pure PMMA
respectively. This is a remarkable achievement considering that these small mass
loadings of the layered hydroxy salt show such noteworthy increments in the thermal
threshold temperature of these composites.
From Table 4.3.1, a general trend is observed were the AT50 values for the CHM
polymer composites are higher than for the ZCM composites. This suggests that the
copper ions in CHM or their oxide forms play an important role in the thermal
stabilization of the polymer matrix. A DTA curve o f the additive ZCM shows several
exothermic processes starting at about 200 C while CHM has one sharp exothermic
feature at approximately 240 C and a broader one at about 350 C. The thermal
combustion of the ZCM additive could therefore increase the amount of heat in the
system rendering PMMA-ZCM less stable in comparison with PMMA-CHM composites.
T50 temperatures for both sets o f PMMA composites fall in the second degradation step,
consisting of chain end and chain scission initiated depolymerization of the remaining

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

polymer. The increments in the threshold temperatures imply that the presence o f the
anionic clays in the polymer matrix severely limits or prevents these depolymerization
processes.

100
80
w
CO

60

cc

40

20
0

50

150

250

350

450

550

Temperature /C
Figure 4.3.9 TGA and DTA curves for pure PMMA, (solid lines) and PMMA-ZCM2%
(bold lines).

Figure 4.3.9 and 4.3.10 shows the TGA and DTA curves for PMMA-ZCM2%
and PMMA-CHM2% composites respectively. TGA and DTA curves for pure PMMA
are also shown to draw a comparison between its thermal degradation behaviors and
those o f the polymer composites. Little change in the shape of the DTA curves of the
PMMA- metal hydroxide composites as compared to pure PMMA is observed. An
endothermic feature in the DTA curve for pure PMMA at approximately 250 C is not
seen in the DTA curve for PMMA-ZCM2%. From Figure 4.3.10, the DTA curve for pure
PMMA and PMMA-CHM2% are similar, showing basically the same features except for
the slight delays seen in the later. An exotherm around 200 C for the PMMA-CHM2%

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

97

composite could be due to the organic fillers contained in the gallery space. From the
collected data we cannot conclusively say whether there is a change in the degradation
mechanism o f the composites relative to the pure PMMA. Future studies involving
monitoring the gaseous products by TGA-FTIR and analysis of residue will afford a
clearer interpretation of the decomposition mechanisms.

100

80
8 60
03
40

20
0

50

150

250

350

450

550

Temperature 1C
Figure 4.3.10 TGA and DTA curves for pure PMMA (solid lines) and PMMA-CHM2%
(bold lines).

The elevation in the T 50 temperature, even without an initial increase in the basal
spacing of the PMMA composites suggests that their thermal stability is a result o f the
interaction of the polymer chains with the residue of additive materials during the
combustion process. The presence of anionic clay modifiers stabilize both steps in the
thermal degradation o f PMMA composites. As the loading percentage of the anionic clay
(ZCM) increased from 2 to 4%, an increase in T 50 o f 9.0 C was observed, with char

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

98

residue also increasing by 3%. Char formed during the thermal degradation, can act as a
barrier to mass and energy transfer hence an increase in the thermal stability of the
composites. This provides direct evidence of the protective role being played by the metal
hydroxide additives.

16000
3s
CO
~

12000

u
0

&

8000

'a>
c
(0

4000
0

30

35

40

45

50

55

60

2 theta (degrees)
Figure 4.3.11 XRD patterns of PMMA-CHM residue from cone calorimetry revealed the
presence o f CuO (), CU2 O (A), and Cu(0) (*) reflections.
PXRD patterns of the residual products revealed the presence of ZnO and CuO
for PMMA-ZCM composites and CuO, Q 12 O, and Cu(0) for the PMMA-CHM
composites and are shown in Figure 4.3.11 and 4.3.12 respectively. However, it is not
clear what role these metal oxides play in improving the thermal stability of the PMMA
composites if any. The reduced Cu(0) and Cu(I) from the combustion of PMMA-CHM
composites imply that there is a role that the Cu2+ ions play in the thermal retardancy
shown by these composites. In spite of the poor dispersion of the anionic clays seen here,
the thermal stability of PMMA-composites obtained is remarkable. This suggests that

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

formation of nanocomposites is not the necessary and adequate cause for improved
thermal stability.

12000
u

9000
4- *

6000

C/3
C
CD

3000

30

35

40

45

50

55

60

2 theta (degrees)

Figure 4.3.12 XRD patterns o f PMMA-ZCM residue from cone calorimetry revealed the
presence of ZnO (0) and CuO ( ) reflections; no reduced species were observed here.
Effects of these additives on PMMA thermal degradation were also measured by
the use of the cone calorimetry with the following parameters obtained including the peak
heat release rate (PHRR), heat release rate (HRR), total heat released (THR), time to self
sustained combustion (TSC), and mass loss rate (MLR). Ideally the favored shifts in the
values of these parameters would be a decrease in the PHRR, time to PHRR, HRR, THR,
and MLR. An increase in the TSC would also be desirable in terms of fire safety o f the
composites.
From Table 4.3.2, the time to sustained combustion is reduced for PMMACHM2% and PMMA-ZCM2%, implying that it is much easier to ignite the composites as
compared to pristine PMMA. There is very little or no reduction in PHRR for all the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

100

modified PMMA composites. Surprisingly significant reductions in the THR are


observed with low modifier loading percentages.
Table 4.3.2

Cone data for pure PMMA and its composites


TSC (s) PHRR (kW/m2) (%red)

Sample

T ph r r ( s)

THR (MJ/m ) (%red)

PMMAa

874

109

69

PMMA-CHM2%

846 (3)

92

53(23)

PMMA-ZCM2%

848 (3)

96

53 (22)

PMMAb

966

104

79

PMMA~CHM4%

902 (7)

87

57 (27)

PMMA-ZCM4%

981 (-2)

83

56 (29)

a> Pure PMMA samples run on different days for different batches.
As much as 29% reduction in the THR is observed for PMMA-ZCM 4%. Polymer
nanocomposites typically show a lower PHRR but a wider heat release distribution
profile with little or change in the THR.10 Organic modifiers bum out at low temperature
producing inorganic metal oxide residues that might slow or inhibit thermal degradation
of the composites.
A reduction in the THR suggests that the amount of material completely burning
out is reduced as we load PMMA with the additives. A 2% CHM loading gave a THR
value o f 53 MJ/m2, corresponding to a 23% improvement in thermal stability with respect
to THR as compared to the virgin polymer. On increasing the loading percent to 4%, the
THR value is reduced by 27% with comparison to pure PMMA. In general, increasing the
loading percent of a fire retardant containing HDS material is expected to further lower
THR value. Figures 4.3.13 and 4.3.14 show plots of the heat release rate as a function of
time for the pure PMMA and its respective HDS and LHS composites at 2% and 4%

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

101

900

700
CM
i

500

_S

300
100

-100
50

100

150

200

time (s)

Figure 4.3.13 Heat release rate curves for PMMA-methacrylate intercalated salt
composites A) PMMA-ZCM2%, B) PMMA-CHM2%, and C) pure PMMA.

900
700
500
300
100
-100
0

50

100

150

200

tim e (s)

Figure 4.3.14 Heat release rate curves for PMMA-methacrylate intercalated salt
composites A) PMMA-ZCM4%, B) PMMA-CHM4%, and C) pure PMMA.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

102

loading respectively. The HRR versus time graphs for the composites show that the
evolution o f heat is spread over a narrow range o f the combustion time. This suggests
that, even though the polymer composites start to bum earlier they extinguish much faster
than the virgin PMMA.
Metal hydroxides introduced into the PMMA are the reason for the improved fire
retardancy. HDSs and LHSs are likely to act as flame-retardants through the three modes
of action. The first mode involves dilution by non-combustible gases (H 2 O and CO2 )
generated from the thermal degradation o f the additives. Oxygen concentration is diluted
by the combustion products mentioned above, snuffing out the flame at its front. In the
second mode, the inorganic combustion products act as condensed phase diluents
reducing combustible matter and hence improving flame retardation. Lastly, Cu may
have some catalytic/chemical role. Hence, fine-tuning the properties o f these layered
materials and their loading percentage has been shown in this work to have a significant
improvement on the thermal stability o f PMMA.

4.4

Conclusions
Even though little evidence is obtained to support the formation of PMMA/HDS

or PMMA/LHS nanocomposites, the results obtained in thermal degradation studies give


hope for a possibility of using these nanostructured materials as additives for improving
the thermal stability o f the polymer composites. The additives have been shown to lead to
no change in PHRR but an intriguing 20-30% reduction in total heat release. TGA
analysis showed a significant improvement in the thermal stability o f the modified
polymers with as high as 61 C increments in selected threshold temperatures for some

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

cases. The improvements obtained point to the fact that the presence of modified
hydroxide inorganic/organic materials in the polymer matrix act as a physical barrier in
shielding PMMA polymer chains against decomposition. Thermal degradation properties
of the composite materials were found to be dependent on the composition o f metals in
the layers and the loading percent of these modified anionic clays into the polymer
matrix.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

104

Chapter 5

5.1

Thermal Degradation Kinetics of Polystyrene (PS) Nanocomposites

Introduction

Polystyrene (PS) has an enormously versatile array of properties and a wide range
of applications in the fast food packaging industry, building insulation, marine, sign
writing, displays and stage propping. In order to improve the physical properties of PS
such as thermal stability, flame retardancy, tensile strength, flexural modulus, reduced
gas and moisture permeability, and solvent resistancy, many researchers have turned their
attention to the formation of polystyrene/modified-montmorillonite (MMT) clay
nanocomposites.166' 178 In nanocomposite formation, modified-MMT clays with
nanometer sized galleries are dispersed into the polymer matrix and the resultant
interfacial interactions between the silicate layers and the polymer matrix is the key to the
observed enhancement in the physical properties. MMT based layered silicates have been
extensively used because of their ubiquity and ease o f processing.
Pristine MMT contains alkali and/ or alkali earth metal ions sandwiched between
negatively charged layers to balance the charge deficiency created by isomorphous
substitution within the layers o f tetrahedral Si4+ by Al3+ or octahedral Al3+ by Mg2+
respectively. Because of the relatively-high cationic exchange capacity (CEC) exhibited
by MMT, it is easy to exchange the incumbent metal ions for ammonium or
phosphonium cations.167' 180The intercalation of long chain ammonium salts results in
increased interlayer spacing and/ or exfoliation, promoting easy penetration of monomer
or polymer chains. The presence o f organic modifiers is vital as it improves compatibility
between the additive and the respective polymer.172 Various methods have been

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

105

employed to prepare MMT based nanocomposites such as melt intercalation,167176,181-183


emulsion polymerization,168 solution blending,171 suspension polymerization,172 and in
situ free radical bulk polymerization.175180
Varying molecular weights are inevitable for all polymer nanocomposites
prepared using the above-described methods, with the exception of melt intercalation
when structurally different additives are used. Wall and coworkers184 and Carasco and
Pages185 have shown that polymer properties, especially thermal degradation, are
dependent on their molecular weights. In our laboratory, polystyrene/modified-MMT
nanocomposites have been prepared using melt intercalation to avoid the problem of
possibly dissimilar molecular weights. The objectives are to (i) confirm the enhancement
in thermal stability following nanocomposite formation, (ii) compare thermal stability of
various nanocomposites as evaluated by thermogravimetric analysis, (TGA), and (iii)
compare their flammability properties using cone calorimetry measurements. From TGA
mass loss derived degradation kinetics of polystyrene nanocomposites a correlation
between observed apparent activation energies and thermal stability, defined here, as the
shift of degradation temperatures to higher values, is examined. Preparation of additives
was done by Grace Chigwada at Marquette University.

5.2

Experiment
Polystyrene (Mwca. 230 000, Mca. 140 000 melt index of 200 C/5Kg), benzoyl

peroxide (BPO) initiator, diethyl ether, tetrahydrofuran (THF), hexadecyl bromide,


acetone, quinoline, and pyridine were obtained from Aldrich Chemical Co.
Montmorillonite (MMT) and Cloisite 10A, an organically-modified montmorillonite,
containing dimethyl, benzyl, hydrogenated tallow, quaternary ammonium (hydrogenated

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

106

tallow is a mixture o f -65% C18, -30% C16, and -5% C14), were kindly provided by
Southern Clay Products, Inc. All chemicals were used without further purification with
the exception of styrene monomer solution where the tert-butylcatechol was removed by
passing through an inhibitor-removal column (Aldrich).
N, iV-dimcthyl-n-hcxadecyl-(4-vinylbcnzyl) ammonium chloride (VB16),166
quinolinium hexadecyl ammonium bromide (QC16),186 and pyridium hexadecyl
ammonium bromide (PYC16)186 were prepared following standard literature methods.
Ammonium cations were exchanged for the Na+ ions in MMT by dissolving these
respective salts in 100 mL o f THF and then mixing with a calculated amount of clay
dispersed in 200 mL of a 2:1 water/THF solution. Based on the cationic exchange
capacity (CEC) of the clay, the amounts of the exchange salts were determined so as to
have a 20% excess. The mixture was stirred at room temperature for 24 h and the
resultant precipitate was continuously washed with water until no halide ions were
detected.
Polystyrene/clay nanocomposites were prepared via melt blending using
established methods.187 Polystyrene nanocomposites were prepared from respective
additives by mixing 5 g of the additive with 95 g of pristine polystyrene to achieve 5%
mass fractions. Melt blending was then performed on a Brabender mixer for 15 min at a
temperature of about 190 C and speed of 60 rpm. The prepared composites are identified
as PS-QC16, PS-PYC16, PS-VB16, and PS-10A for QC16, PYC16, VB16, and 10A
additives respectively. A reference sample o f pure polystyrene was obtained by following
the same procedure only without an additive. X-ray diffraction (XRD) patterns, TEM

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

107

images (David Jiang at Cornell), TGA, and cone calorimetry measurements were
obtained as described in Chapter 4.

5.3

Results and Discussion


Closite 10A (10A), pyridium hexadecyl ammonium bromide (PYC16),

quinolinium hexadecyl ammonium bromide (QC16), and A, A-dimethyl--hexadecyl-(4vinylbenzyl) ammonium chloride (VB16) were used to modify the Na+ intercalated MMT
and their chemical structures are shown in Figure 5.3.1. These surfactants were carefully
chosen to demonstrate particular points of interest in nanocomposite formation. All salts
contain a long organic chain (C l6) which is chemically attached to different moieties
such as pyridinium, quinolinium, styrenic, and a phenyl ring. We seek to investigate the
effect of the end groups on the thermal stabilization o f PS nanocomposites. Comparative
studies will be done between current and previous work where different preparative
methods were used.

PYC16

QC16

VB16

10A

Figure 5.3.1 Structures of the salts used to prepare the organically modified clays.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

108

4500

12000

PS-10A

9000

-4'

c
3

o
o

PS-VB16

6000

to

2 thetaidegrees)

CD

3000

C lo is ite 10A

V B 16

10

15

20

25

30

2 theta (degrees)
Figure 5.3.2 XRD data for Cloisite 10A and VB16 clays. Insert shows XRD data for
Cloisite 10A and VB16 polystyrene/clay nanocomposites. Data are offset for clarity but
otherwise not scaled.
The microstructures of modified-MMT clays, QC16 and PYC16 and their
corresponding polystyrene nanocomposites were examined by XRD and have been
reported by the Wilkie group.186 Exchanging Na+ ions with organic cations resulted in
significant increments in the basal spacing. The characteristic 001 peak for Na+
containing MMT is at 20 = 7.9, corresponding to basal spacing, d, o f 11.2 A while the
007 peaks for the modified clays are shifted to lower 20 values (e.g. larger d-spacing
values) confirming a successful intercalation of the longer cations. The XRD patterns for
10A and VB16 clays and their respective PS nanocomposites are shown in Figure 5.3.2.
Basal spacing increments of 6.7 and 21.0 A were observed for 10A and VB16 following
exchange of Na+ ions respectively. Following melt intercalation, 00/ reflections were
further shifted to lower 20 values in most cases, indicating that polymer chains penetrated
into the galleries, consequently enlarging the interlayer space between the clay layers.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

109

The calculated basal spacings are 34.0, 29.0,186186 33.0,86186 and 24.3A for PS-10A, PSQC16, PS-PYC16, and PS-VB16 polystyrene/clay nanocomposites respectively. We
however note here a reduction in the basal spacing o f PS-VB16 when compared to the
interlayer spacing observed in VB16. This is not uncommon as Jang and coworkers188
made a similar observation for polystyrene nanocomposites made via melt intercalation.
The 001 peak is very broad suggesting a substantial amount o f layer disorder and
possibly exfoliation. The intercalated polymer chains are known to adopt a flattened
configuration between the silicate layers, since the melt radius of gyration of PS chains (~
95.9 A)183189is larger than the observed increments in d-spacing upon melt intercalation
in all cases.
From XRD analysis, the intercalated MMT phases in the nanocomposites are
obvious. However XRD only shows periodic stacking of layers. Other methods like
transmission electron microscopy (TEM) can be employed to directly investigate whether
some exfoliated phases exist within the polymer matrix. The low magnification images
provide information about the nanodispersion while high magnification images tell
whether exfoliation and/ or intercalation has been achieved. Chigwada and coworkers186
reported the TEM image o f PSQC-16 at low resolution to appear as regions o f alternating
narrow, dark bands, and wide light bands. The dark lines correspond to clustered clay
particles (tactoids and agglomerates). The absence o f parallel and equally spaced lines at
higher resolution suggests no intercalation of PS into QC16 layers but this is not
consistent with the XRD results. They also observed that the dispersion of clay layers is
improved for the PS-PYC16 as can be seen in the low magnification image. The higher

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

110

magnification image permits the observation of discrete clay layers suggesting co


existence of exfoliation and intercalation.
The TEM image o f PS-10A at higher magnification is shown in Figure 5.3.3. The
10A layers in parallel registry (dark regions) reveal periodical stacking consistent with
intercalation as evident from XRD results. Higher resolution TEM image of PS-VB16 is
shown in Figure 5.3.4. The overall picture shows that the modified MMT layers did not

Figure 5.3.3 TEM image at high magnification for PS-10A. The scale bar (bottom centre)
represents 100 nm.

a
Figure 5.3.4 TEM image at high magnification for PS-VB16. The scale bar (bottom
centre) represents 100 nm.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Ill

occupy the full volume with regions o f pure PS visible. The microscopic image suggests
existence of both intercalation and exfoliation. Zhu and coworkers showed a totally
intercalated PS-VB16 nanocomposites using bulk polymerization . 166 Gilman and
Morgan190 reported an exfoliated PS nanocomposite using VB16 using the same method
We note here that different preparative methods lead to different additive distribution and
behavior within the polymer matrix. However, in all cases, the presence o f a fairly good
homogeneous distribution of respective additives shows that the melt intercalation
method generally has favorable thermodynamics of mixing.
Thermogravimetric analysis (TGA) decomposition behaviors of pure polystyrene
and its nanocomposites are shown in Figure 5.3.5A while their corresponding derivatives
(DTG) are shown in Figure 5.3.5B. The onset of thermal degradation, measured as the
temperature at which

10

% mass loss occurs, Ti0, is shifted to higher temperatures for all

nanocomposites relative to pure PS. This behavior is maintained throughout the thermal
decomposition as clearly evident from the shift of T5 0 , the temperature at which 50%
mass loss occurs, and Tmax, temperature at maximum degradation rate, to higher values;
in some cases AT5 0 (T50 for PS nanocomposites minus T 50 for PS) is as high as 22 C.
Given the small amount of the clay phase, this improvement in thermal stability is
significant. From DTG curves, Figure 5.3.5B, meaningful mass losses are seen from
about 375 C, suggesting that the modified clays are thermally stable to such high
temperatures; we are confident the melt blending process at 190 C did not perturb the
structural integrity of the respective additives. These data are summarized in Table 5.3.1.
Other authors have reported similar observations for polystyrene nanocomposites
prepared by melt blending using modified MMT clays. 1 7 4 ,191 TGA results for the same

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

112

system might be different in the work reported186 earlier because different TGA
instruments were used.

PS
O

P S -1 0 A
P S-V B 16
P S -Q C 1 6

P S -P Y C 1 6

PS
o

P S-10A
PS-V B 16
P S -Q C 1 6

300

350

400
450
500
Temperature /C

P S -P Y C 1 6

550

600

Figure 5.3.5 (A) TGA curves for pure PS (dashed line), PS-VB16 (solid line), PS-PYC16
(solid diamonds), PS-10A (empty circles), and PS-QC16 (bold line). (B) DTG curves for
pure PS (dashed line), PS-VB16 (solid line), PS-PYC16 (solid diamonds), PS-10A
(empty circles), and PS-QC16 (bold line). Derivatized mass losses are scaled by a factor
of 100.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

113

Table 5.3.1

TGA data for polystyrene nanocomposites


Tio(C)
3970
4051
4091
4071
4091

/s
O
O
O
H

Sample
Pure PS
PS-10A
PS-QC16
PS-VB16
PS-PYC16

4260
4420
4481
4390
4420

AT5 0 (C) Tm(C) Char (%)


4311
0
1 0 (0)
160
4451
31 (3)
221
4500
41 (4)
130
4401
40 (3)
160
4430
31 (4)

Tio, temperature at which 1 0 % mass loss occurs; T5 0 , temperature at which 50% mass
loss occurs; AT5 0 , T 50 (composite) minus T50 (pure PS); Tm, temperature at maximum
degradation rate. Italicized entries are the expected char based on the residue obtained
from pure PS and additive fractions.

Mass difference curves (mass % o f PS nanocomposites minus mass % of pure PS


at the same temperature) for polystyrene nanocomposites are shown in Figure 5.3.6 A. A
mass percentages are positive over the entire degradation temperature range for all
investigated nanocomposites. This shows that the polystyrene nanocomposites are more
thermally stable than the virgin polymer under the thermo-degradation processes
involved. Maximum stabilization was observed at the temperature at which the highest
mass loss rate occurs, an indication that addition of various MMT based additives
prevents or retards depolymerization occurring via random chain scission. Notably, the
amount of char remaining after complete combustion is higher for the nanocomposites
when compared to pure PS. The accumulation of a char provides a physical barrier that
stops the evaporation of small molecules generated during the thermal decomposition,
effectively restricting uninterrupted decomposition o f the PS.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

114

O P S -1 OA
P S -V B 1 6
P S -Q C 1 6

50

150

P S -P Y C 1 6

PS
PS-10A
PS-VB16
PS-QC16
PS-PYC16

250
350
450
Temperature /degC

550

Figure 5.3.6 (A) Curves o f mass loss differences for PS nanocomposites; pure PS (dashed
line), PS-VB16 (solid line), PS-PYC16 (solid diamonds), PS-10A (empty circles), and
PS-QC16 (bold line) as a function of degradation temperature. (B) DTA curves for pure
PS (dashed line), PS-VB16 (solid line), PS-PYC16 (solid diamonds), PS-10A (empty
circles), and PS-QC16 (bold line).

Also the presence of clay platelets confines PS chains and the restricted thermal motion
serve to insulate them from heat, therefore char formation provides a physical barrier to
both mass transport and heat transfer. However, the proposed thermal insulation is

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

115

dependent on the identity of intercalates and perhaps their percent constituents since all
four additives stabilize PS to different extents even though they are all MMT based. A
simple comparison of the observed shifts in thermal degradation temperatures suggests
the following order regarding thermal stability for the nanocomposites under
investigation: PS-QC16 PS-PYC16 ~ PS-10A > PS-VB16. The presence o f a
quinoline group in QC16 may render it more effective in thermal stabilization when
compared to other additives containing phenyl groups.
Differential thermal analysis (DTA) curves of pure PS and its nanocomposites are
shown in Figure 5.3.6B. The DTA curve for pure PS is similar to that of its respective
nanocomposites with the exception that the endothermic feature at around 440 C is
shifted to relatively higher temperatures and is less pronounced for the nanocomposites.
This is additional evidence of enhanced thermal stability for the nanocomposites as
manifested in the positive degradation temperature shifts. Since the DTA profiles of
nanocomposites have the same features as pure PS, it is proposed here that there is no
significant change in the degradation mechanism, however, the degradation processes are
delayed. To elucidate mechanistic changes if and/ or when they occur, evolved products
were both qualitatively and quantitatively monitored using gas chromatography-mass
spectroscopy (GC-MS) and thermal gravimetric analysis-Fourier transform infrared
(TGA-FTIR) for a PS-10A loaded at various percentages.192 It was proposed that the
degradation pathway of polymers was modified by the presence of the clay. Similarly,
identification of evolved gas products and the inorganic residues from the systems being
investigated here are necessary to detect mechanistic changes if they occur. This will be
the subject of future studies.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Thermal degradation o f pure polystyrene has been shown to occur in a single step
with styrene monomer and oligomers as the primary volatile products, together with
smaller amounts of benzene and toluene (combustibles).193 The high degree of accuracy
with which mass measurements are done in TGA experiments provides a means to
extract reliable and precise estimates o f kinetic parameters such as activation energies,
frequency factor, rate o f decomposition, and in some cases the order of the
decomposition reaction. Both isothermal and non-isothermal methods can be used to
estimate the above-mentioned parameters. In this study a non-isothermal method was
used rather than a conversional isothermal one for the following reasons; (i) it is possible
to retrieve a characteristic information for a given sample over the entire degradation
temperature range, (ii) samples might undergo some side reactions during the process of
raising the temperature to the desired isothermal value, and (iii) kinetics are established in
a continuous fashion over the entire decomposition profile. Dynamic thermogravimetric
analysis (linear temperature increase) was employed in this study using multiple heating
rates kinetics to estimate apparent Ea values for the polystyrene nanocomposites.
There are several mathematical models that can be used to obtain kinetic
parameters from these solid-state kinetic data and these are conveniently divided into two
groups namely model-fitting and isoconversional (model-free) methods. Model fitting
methods have been extensively criticized for non-isothermal applications as regression
methods used with these methods sometimes lead to indistinguishable fits between
different models.194' 197 Another major disadvantage inherent in model-fitting methods is
that they yield a single average value for the entire degradation process, while the thermal
degradation o f polymeric materials normally involve multiple steps characterized by

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

117

different activation energies.197 Isoconversional methods have gained wide spread use as
they do not have modelistic assumptions,197'200 however, their use is governed by
assumptions that they are independent o f the dynamic heating rate or temperature. It is
therefore possible to calculate the apparent activation energy, Ea, as a function of the
degree of conversion, a.
The multiple heating rate kinetics (MHRK) method is used in this study to extract
effective kinetic parameters from these rather complicated decomposition processes,
applying the frequently used Flynn-Wall-Ozawa201 method specifically derived for
AAA

heterogeneous chemical reactions under linear heating rates.

The derivation of the

Flynn-Wall-Ozawa method from first principles is presented in reference 201. Simplified,


the Flynn-Wall-Ozawa method is expressed by the equation:
log f (a) = log (AE a /R) - log (p) - 2.315 - 0.4567E /RT

5.3.1

where f (a) is known as the conversional functional relationship; A is the pre-exponential


factor (min'1), E a is the apparent activation energy (kJ mol'1), R is the gas constant, p is
the heating rate (C /min), and T is the absolute temperature (K). Equation 1 can be
further simplified to
log (P) = - 0.4567 E a /RT + constant

5.3.2

Activation energies were calculated from the slopes o f the isoconversional plots of log
(P) vs. 1/T for fractional conversions a = 0.10 - 0.90 at intervals o f 0.05. These activation
values are plotted as function o f a in Figure 5.3.7 below.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

118

270 Pure PS
PS-PYC16
PS-QC16
PS-VB16
APS-10A

250
C 230
o

* 210

A 11
lud

O !>

190 170 150 0

0 .2

0 .4

0 .6

0 .8

C on version Fraction (a)

Figure 5.3.7 Plot o f Ea versus extent o f decomposition, a, for the non-isothermal


degradation o f pure PS (empty circles), PS-VB16 (solid triangles), PS-QC16 (empty
diamonds), PS-PYC16 (solid cirles), and PS-10A (empty squares).

From Figure 5.3.7, the initial activation energy at 10% conversion signaling the
onset o f the decomposition process for pure polystyrene was 170 kJ m ol'1. At about 30%
conversion, the E a values has risen to -190 kJ m ol'1and this activation value is somewhat
maintained throughout the degradation profile. Comparable results were reported by
I Q O

Vyazovkin and coworkers

using commercial polystyrene o f molecular weight 124, 000

and polydispersity index o f 2.7. The activation energy o f PS-QC16 is the highest (260 kJ
mol"1) followed by that of PS-PYC16 (220 kJ m ol'1) at a = 0.10 consistent with the
observed slight delay in the onset o f thermal degradation, (Tio values, Table 5.3.1) for
these composites relative to pure PS, PS-10A, and PS-VB16. However, the Ea values for
PS-QC16 decrease with conversion to reach a constant value o f 215 kJ m ol'1 in a range

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

119

of 0.40 - 0.90. A similar trend is observed for PS-PYC16 where the Ea decreases from a
maximum value o f 220 kJ mol'1 at a = 0.20 to a constant value o f 210 kJ mol'1 in the a
range of 0.40 - 0.90. The activation energies o f PS-10A and PS-VB16 rapidly increase to
maximum Ea values of 230 and 220 kJ mol'1 at a = 0.30 respectively before gradually
falling to an average Ea value of 210 kJ mol'1.
Overall, the Ea values for the nanocomposites are higher than for pure PS in the
fraction conversion region o f 0.40-0.90 by ~ 20 kJ mol'1. This is consistent with the
observed enhanced thermal stability for the nanocomposites marked by the shift to higher
temperatures o f their TGA curves (measured at a constant ramp rate; 20 C/min) relative
to pure PS. However, based on the activation energy trends for the nanocomposites
studied herein, it is hard to discern their order of thermal stability. The fact that the Ea
values for the polystyrene nanocomposites are similar within experimental error in the a
range of 0.4 - 0.9 suggests that there is not a large shift in the degradation mechanism
between the different clays at higher conversion fractions. A wide variation in activation
energies is, however, observed for low conversion fractions and this may be attributed to
the varying thermal stability o f the surfactants in the low temperature regime.
Initiation o f polymer degradation is attributed to the existence o f weak links, such
as head-to-head,204 hydroperoxy, and peroxy193 structures, within the polymer
composites, which easily yield at very low temperatures to form radicals that participate
in subsequent degradation process at elevated temperatures. The increase in activation
energies at lower conversion fractions (a < 0.30) suggests a shift o f the rate-limiting step
from disintegration of these weak links to mainly random scission. Most researchers have
attributed the improved thermal stability and fire retardancy in polymer/clay

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

nanocomposites to the existence o f a physical barrier limiting mass transport and energy
transfer. Thermal stabilization of polystyrene at conversion fractions o f less than 0.30,
where no meaningful char formation would have occurred, suggests that the chemical
effect of clays contribute immensely to the observed property enhancements. However, at
higher degradation temperatures char formation may greatly contribute to the observed
enhanced thermal stability via the cage effect phenomenon where the gaseous
molecules formed during thermal decomposition are trapped within the solid inorganicorganic network. The complexity o f decomposition reactions of polymeric materials
precludes the authors from ascribing the stabilization effect at high degradation
temperatures to char formation alone; several factors, including chemical reactions, might
be at play as well.
The heat release rate (HRR) as a function of time for pure PS and its composites
at a flux of 35 kW/m2 are shown in Figure 5.3.8. A slight reduction in the time to
sustained combustion is observed for all the nanocomposites except for PS-QC16, where
TSC remains the same compared to the virgin polymer. This correlates with the observed
delay in the onset of thermal degradation seen in the TGA profile and the elevated
activation energy observed for PS-QC16 when compared to other composites. The HRR
curves for the nanocomposites show that the evolution o f heat is spread over a similar
range of the combustion time as pure PS with the exception of PS-QC16 showing a
prolonged combustion time. PS-PYC16 did not show any reduction in PHRR while PSQC16 gave the greatest reduction, (34%). A significant reduction in AMLR (~ 20%) was
observed for PS-QC16 suggesting that a smaller amount of heat is released per unit time
hence the reduction in PHRR.

T PH r r

was prolonged for all the nanocomposites relative to

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

121

PS
o

1180

P S -1 0 A
P S -V B 1 6
P S -Q C 1 6

CM

P S -P Y C 1 6

<J5

7 80

O*

380

>
o

-20
0

50

100

150
Time (s)

200

250

Figure 5.3.8 (A) Heat release rate curves pure PS (dashed line), PS-VB16 (solid line), PSPYC16 (filled diamonds), PS-10A (empty circles), and PS-QC16 (bold line) from cone
calorimetry measurements at 35 kW/m . Curves for PS-PYC16 and PS-QC16 are as
reported in reference 186.

pure PS, with PS-QC16 showing the greatest delay o f ~ 20%. Addition of clay promotes
char formation205 which heavily retards diffusion o f combustible products from low
temperature regime polymer pyrolysis hence the significant reduction in heat release rates
for nanocomposites relative to the pure PS. Table 5.3.2 gives a summary of the results
obtained for PS and its nanocomposites with different additives.
Zhu and coworkers166 prepared VB-16 containing polystyrene nanocomposites at
various loadings via bulk polymerization. The TSC for the polystyrene nanocomposite
was the same as for the pure PS unlike for the melt blending sample were a significant
reduction is observed for the nanocomposite. More than two fold improvements in
PHRR, THR, AMLR, and ASEA were observed with bulk polymerized samples.166 This
shows that the fire properties of polymer nanocomposites produced using different

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

122

preparative methods are dissimilar. In making this proposition we assume the molecular
weight distributions of the pure polystyrene and polystyrene nanocomposite are the same.

Table 5.2.2

Cone calorimetry data for polystyrene nanocomposites at a flux o f 35


kW/m2.

AMLR

ASEA

(g/stm2)

(m2/kg)

1138

THR
(MJ/m2)
(red%)
999

301

115950

1351 87

12621

1001

311

126523

483

108125 (16)

120 7

912 (8)

271 (10)

118714

PS-QC16

602

84829 (37)

138 6

943 (6)

241 (22)

132824

PS-VB16

432

101725 (21)

12912

867(13) 272 (10)

115023

PS-PYC16

492

131977 ( 2)

126 7

973 (3)

140833

TSC

PHRR (kW/m2)

T phrr

(s)

(red%)

(s)

PS

594

1284180

PSa

634

PS-10A

Sample

302 (3)

TSC, time to sustained ignition; PHHR, peak heat release rate; T p h h r , time to peak heat
release rate; THR, total heat release; AMLR, average mass loss rate; CY, char %; ASEA,
average specific extinction area (a measure of smoke). a Previous work published
elsewhere; reference 186.

However, the work reported does not mentioned molecular weight distributions
for obtained samples. Jang and coworkers192 reported a 55% reduction for a 5% loaded
bulk polymerized polystyrene/10A nanocomposite. In this work we observed 16%
reduction in PHRR. Gilman and coworkers 168 suggested that significant reductions in the
PHRR are a result of good nanodispersion of the additive within the polymer matrix.
Even though TEM images show fairly good additive distribution in the polystyrene
nanocomposites, cone calorimetry results suggest that the dispersion is not enough to
cause significant reductions in the PHRR.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

No significant percent reductions in the total heat released are observed for the
nanocomposites under investigation. This suggests that even though a protective barrier is
formed at the surface of the polymeric material its efficiency is not sufficient to
effectively inhibit combustion. The formation of ceramic surfaces during combustion o f
polymeric material may serve to slow down the flow of combustibles, thus successfully
reducing the flame intensity; however, without a significant reduction in the amount of
polymeric material burning out at the conclusion of the combustion process. The
presence of the inorganic/organic modifiers in the PS nanocomposites may provide heat
sinks during combustion, perhaps via efficient thermal transfer of energy to metal
components of the additives hence protecting polystyrene, which has a low thermal
conductivity. No significant reductions in the smoke content are observed for the
polystyrene nanocomposites relative to the pure PS. The improvement in fire retardancy
is shown by the reduction in the PHRR, reduction in AMLR in some cases, the increase
in the

T phrr,

and char remaining after complete combustion.

Kinetic information used to estimate the activation energies of polystyrene and its
nanocomposites is derived solely from mass loss profiles of thermal degradation
processes. The mathematical function used to describe the degradation process is
modeled for the released products, i.e. from weight loss, and provides no information
about which, or how many bonds are broken at any given conversion fraction.195 The
extracted apparent Ea values are thus useful here to explain the shift to higher threshold
temperatures for PS nanocomposites but not necessarily their flammability. It is
imperative to recognize that kinetic analysis does not take into account mass transport
and heat transfer during thermal decomposition o f polymeric materials, thus no

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

124

correlation can be expected between activation energy trends and flammability property
measurements from cone calorimetry.
For uninterrupted burning o f polymeric material to occur, the initial heat must be
adequate to cause decomposition o f the material, the temperature must be high enough to
ignite the products of thermal degradation, and the heat transferred back to the polymer
must be sufficient to maintain the cycle. An effective fire retardant must break this cycle
somehow. This can be achieved by (i) enhancement in thermal stability o f the polymeric
material; (ii) quenching of the flame; and (iii) retardation of heat supply. MMT based
flame-retardants may operate via several modes of action: (a) sacrificial absorption of
heat by metallic Mg or A1 (heat sink) from the intralayer structure of montmorillonite
preventing heat transfer to polymeric material thus improving thermal stability; (b)
dilution of combustion atmosphere by non-combustible gases (H2 O and CO2 ) generated
from the thermal degradation of the additives (flame quenching); (c) the inorganic
residual products may act as a physical barrier (cross-linked char) reducing diffusion of
combustible matter (mass transport) and/ or energy transfer, hence improving flame
retardation. The success of a fire retardant depends on at least one, some or all o f the
modes of action described above.

5.4

Conclusions
The thermal stability, degradation kinetics, and combustion behaviors of PS and

its nanocomposites have been studied using both TGA and cone calorimetry. Notable
char formation suggests the effectiveness of these additives as potential fire retardants. In
TGA experiments, significant increments in T 10 , T 50, and Tmax are observed for all

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

additives under investigation. Addition o f small amounts of the modified-MMT clays


(5%) to pure PS result in significant reductions in the peak heat release rate (PHRR) as
measured using cone calorimetry with the exception o f the PS-PYC16 nanocomposite.
The average mass loss rate is reduced by 20% for PS-QC16 and this is consistent with the
observed enhanced thermal stability measured from TGA experiments. However, there is
no significant reduction in total heat release (THR). The presence o f the organically
modified clays during the course o f degradation of the nanocomposites is likely to
catalyze the formation of cross-linked char resulting in effective thermal stabilization of
these polymeric materials.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

126

Chapter 6

Nanostructured Layered Copper Hydroxy Dodecyl Sulfate:


A Potential Fire Retardant for Poly(vinyl ester) (PVE)

6.1

Introduction
Thermal stability of polymers has been enhanced by blending virgin polymers

with additives such as particulate inorganic oxides, like alumina trihydrate (ATH) and
magnesium hydroxide (Mg(OH)2 ).

0 ( \f\ 0(Y 7

However, the disadvantage o f ATH and

Mg(OH ) 2 is that they are most effective at very high loadings, about 65%, which can
have detrimental effects on the mechanical properties of the composite. Halogencontaining fire retardants (FR) have been used in engineering thermoplastics and epoxy
resins to improve their thermal stability and fire performance.206'209 Despite their
demonstrated effectiveness in reducing flammability, the use of halogen-containing fire
retardants in commercial plastics is limited because of their corrosivity and potential
toxicity 210,211 Various non-halogen containing fire retardants, such as metal oxides,
metal hydroxides, metal salts, nitrogen-containing, phosphorus-containing and cellulose
fibers, have been used to enhance the thermal behavior of polymers.212'214 Phosphoruscontaining additives have shown excellent thermal stabilization effects on polymers,
however they tend to cause plasticization.215 It has been postulated that phosphates are
oxidized to phosphoric acids during combustion and these acids may alter the degradation
pathways of the polymer and promote char formation.206
Polymer/clay nanocomposites containing smectite clays such as montmorillonite
(MMT), hectorite (HET), and magadiite (MGH) have been shown to provide
enhancements in physical properties o f polymers such as increased tensile strength,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

127

tensile modulus, flexural strength, thermal stability, corrosion protection and reduced
permeability and moisture absorption, owing to their structural morphology.61'67 Recently,
several classes o f synthetic inorganic/organic compounds structurally similar to natural
fl

smectite clays have emerged as potential fire retardants. ' These include layered
double hydroxides (LDHs), hydroxy double salts (HDSs), and layered hydroxy salts
(LHSs), all containing positively charged metal hydroxyl layers with negatively charged
anions contained in nanodimensional galleries. The ability to vary the identities of metals
and interlayer anions provides design parameters that can be tuned to optimize additive
effectiveness.
Previous work has shown a significant reduction (~20 to 30%) in total heat
release (THR) using cone calorimetry when copper-containing hybrid inorganic/organic
layered compounds, copper hydroxy methacrylate (CHM) and zinc/copper hydroxy
double salts were used with polystyrene (PS) and poly(methyl methacrylate) (PMMA).52,
54,216 In the work reported herein, an LHS additive, copper hydroxy dodecyl sulfate
(CHDS) is added to PVE to explore its potential utility as a fire retardant. A long organic
chain the size o f dodecyl sulfate is used in order to improve the hydrophobicity of the
interlayer space.
Analysis of residues by x-ray diffraction (XRD) and Fourier transform infrared
(FTIR) spectroscopy provides evidence of the role of chemical effects, due to the
presence of copper, in stabilizing the composite. Various mechanisms explaining
enhancements seen in polymer thermal and flame stabilities on addition of fire retardants
are proposed and correlation with experimental observations is attempted.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

128

6.2

Experimental
The chemicals used in this study: vinyl ester resin, bisphenol-A/novalac epoxy,

mass fraction of 67% in styrene, [Derakane 441-400] (Ashland Chemical Company); the
initiator, 2-butanone peroxide [BuPO]; cobalt naphthenate catalyst [CoNp] (Aldrich
Chemical Co.); sodium dodecyl sulfate, (75.0%) [SDS]; FTIR grade-potassium bromide
[KBr] (Alfa Aesar); hydrated copper nitrate, (98.9%) [Cu(N0 3 )2 .2 1/ 2 H 2 0 ] (Fisher
Scientific Company); and aqueous ammonium [NH4 OH] (EM Science, Merck), were
used as received.
A layered hydroxy salt, (LHS) copper hydroxy nitrate, Cu2 (0 H)3 N 0 3 , (CHN) was
prepared via a standard literature method .47 Copper (II) nitrate, (100 g; 0.430 mol) was
added to 1 L of distilled water and the pH o f the resultant solution adjusted to 8.0 0.1
by addition of aqueous ammonia. The dispersion was aged for 24 h after which the
precipitate was filtered off, washed, and dried. Dodecyl sulfate anions were exchanged
for the NO 3 anions in CHN by mixing the dried CHN precursor material with 0.2 M
solution of sodium dodecyl sulfate. In the anion exchange process, 10 g of CHN were
mixed with 500 mL of the exchange solution and frequently shaken for 48 h. The clear
supernatant was decanted and replaced with a fresh solution of SDS for an additional 48
h, after which the exchanged product, copper hydroxy dodecyl sulfate,
Cu2 (0 H)3 (CH3 (CH2 )ii 0 S0 3 ), (CHDS), was recovered by filtration, washed, and dried.
Vinyl ester composites (~ 120 g) were prepared at room temperature by mixing
the vinyl ester resin and fire retardant(s) using a mechanical stirrer for 3 h. The initiator,
BuPO (1.3%), was then added and the mixture stirred for a few minutes, followed by
addition of the catalyst, CoNP (0.3%), and stirring continued for a few minutes to achieve

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

129

homogeneity. Pure PVE was loaded to afford n% fraction of the additive, CHDS, within
the polymer matrix, yielding composites identified as PVE/CHDS-. Percent loadings
were determined from the final mass of the composite, assuming no loss of the additive
during the preparation process. Approximately 30 g samples were rapidly transferred to
pre-formed 10 cm x 10 cm x 3 mm aluminum dishes, making platelets of uniform
thickness for cone analysis. Samples were allowed to cure overnight at room temperature
and then post-cured at 80 C for 12 h.
X-ray diffraction (XRD) patterns of the synthesized layered materials and PVE
composites were obtained as previously discussed in chapters 2 and 4. Polymer
composite samples were poured onto aluminum foil to make approximately 1 mm thick
platelets, which were then cured in the same manner as samples for cone calorimetry. The
resultant platelets were then mounted onto vertically oriented sample holders for XRD
analysis. Assignments of phases o f known copper containing species were made using
the powder diffraction data base.147
Fourier transform infrared (FTIR) spectra, TGA analysis, and cone calorimetry
for the solid materials and composites were obtained as described in chapters 2 and 4.
Bright field transmission electron microscopy (TEM) images were collected at 60 kV
using a Zeiss 10 c electron microscope by David Jiang at Cornell University.

6.3

Results and Discussion

Figure 6.3.1 shows the chemical structures of bisphenol-A and novalac


epoxy vinyl ester resins. Radicals generated by the decomposition o f the initiator, BuPO,
initiate radical polymerization. Since the novalac epoxy vinyl resin has several pendant

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

130

Bisphenol-A epoxy vinyl ester

OH

OH

Novalac epoxy vinyl ester

n(l-6)

Figure 6.3.1 Chemical structures of bisphenol-A and novalac epoxy vinyl resins.

polymerizable vinyl containing groups, the resultant PVE will be highly cross-linked. Xray diffraction (XRD) is commonly used to characterize the interlayer spacing of layered
inorganic-organic hybrids. Figure 6.3.2 shows the powder x-ray diffraction pattern of the
layered hydroxy dodecyl sulfate, CHDS additive. The samples were partially exchanged
as evident from reflections due to the precursor, copper hydroxy nitrate (PDF#14-687)147
which are marked with solid triangles. The structure of CHN is shown in Figure 6.3.3a.
Two distinct phases of the exchanged product, marked with asterisks and solid
circles are observed. The average basal spacing, d, for the phase marked with solid circles
was found to be 25.9 0.9 A. This suggests that the dodecyl sulfate (DS) anions are
arranged perpendicular to the hydroxylated metal sheets forming a monolayer array
within the gallery space. The sulfate ends of the DS are bound to the positively charged
LHS layers in an interdigitated fashion, maximizing the van der Waals attraction as

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

131

16000
"

3 " 12000
8000

S ' 12000 03-

c j PVE/CHDS- 3

4000

b) PVE/CH DS-1

a) CHDS (x %)

8000 -

to
c

theta (degrees)

4000 -

10

20
30
40
2 theta (degrees)

50

60

Figure 6.3.2 XRD data for partially exchanged CHDS reveal the presence of two new
phases; (*) with basal spacing, d= 39.2 A and () with d= 25.9 A. Reflections from the
precursor, CHN, are present (A). The insert shows the XRD patterns of the PVE-n
composites at n = 1, 3, 5, 7, and 10% loadings, labeled PVE/CHDS-1, PVE/CHDS-3,
PVE/CHDS-5, PVE/CHDS-7, and PVE/CHDS-10 respectively and also the XRD pattern
of the partially exchanged CHDS, scaled by a factor o f %. Data for the PVE composites
are offset for clarity but not otherwise scaled.

shown in Figure 6.3.3b. Several researchers have reported similar results with layered
double hydroxides (LDHs), which are inorganic/organic layered hybrids resembling the
LHSs.100,217The second LHS-DS phase, marked with asterisks in Figure 6.3.2, was
calculated using an average of 00/ (/ = 1 and 2) reflections to have a basal spacing, d, o f
39.2 0.2 A, consistent with the DS anions arranged either in a partially interdigitated
bilayer or tilted in such a way to have little or no overlap. Figure 6.3.3c shows a bilayer
arrangement with DS anions arranged in a perpendicular fashion to hydroxy layers with
no overlap. This phase will be designated as the bilayer orientation.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Figure 6.3.3 Schematic representations of (a) the layered copper hydroxy nitrate (CHN);
copper (grey), nitrogen (black), and O/OH (white), (b) monolayer, and (c) bilayer phases
of CHDS.

Formation of two new exchanged phases is not uncommon for the anionic
exchange in layered hybrid inorganic/organic copper (II) compounds.20 This has been
attributed to the coexistence of both gauche and trans arrangements of the exchanging
anions in the interlayer spacing or possibly the presence of phases with different degrees
of hydration. In the case of dodecyl sulfate, it is most likely that differences in orientation

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

of the alkyl chains leads to the formation of the two phases. The overall extent o f anion
exchange has been found to be dependent on the chain length. In the case of zinc copper
acetate hydroxy double salt, exchange with octanoate at room temperature proceeded to
completion while shorter exchange anions such as butyrate did not completely exchange,
chapter 2. This is consistent with a thermodynamic stabilization effect for the longer alkyl
chains.30Park and Lee have recently reported218 alkyl chain length effects in CHN
exchanged with sulfonates. In the case of CnH2 n+iS0 3 ' anions with n = 12-18, incomplete
exchanges were observed while those with n = 6-10 proceeded to complete exchange, an
effect attributed to the weak coordinating power o f the sulfonate.
In the insert in Figure 6.3.2, the XRD patterns of the polymer/CHDS composites
at a range o f loadings from 1 to 10% are shown. Features due to the additive are evident
at higher loadings. All phases from the CHDS except the one with a 39.2 A d spacing are
retained in the composites, without displacement from their original 20 positions. This
suggests that no polymer chains and/ or monomers were intercalated into the gallery
spaces of either the CHN or the monolayer phase o f CHDS (d = 25.9 A). However, the
disappearance of the CHDS reflections at 20 values o f 2.3 and 4.5 (bilayer phase; d =
39.2 A) from the XRD patterns of the composites even at 10% loadings suggests the
formation of either an exfoliated andJ or intercalated nanocomposite. Exfoliation would
lead to a disordered arrangement of LHS layers, hence the disappearance of reflections at
20 values of 2.3 and 4.5 . Intercalation of polymer chains could result in a very large d
spacing corresponding to very low 20 values, which would be beyond the detection limit
of the wide-angle x-ray diffraction apparatus used in this study. The possibility of free
spaces within the interlayer domain of the bilayer phase suggests that monomeric units

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

134

and/ or polymer chains can be incorporated hence the possibility of exfoliation and/ or
intercalation.

Figure 6.3.4 TEM images at low (left) and high (right) magnification for PVE/CHDS-5.
For low magnification, the scale bar (bottom center) represents 500 nm while for high
magnification, the scale bar (bottom center) represents 100 nm. Regularly spaced stripes
running diagonally across the image are an artifact due to sample preparation.

XRD alone can, in some cases, demonstrate the formation o f intercalated


nanocomposites, but cannot distinguish between exfoliation and immiscibility.
Transmission electron microscopy (TEM) becomes an important complementary tool to
investigate the morphology of the composites. Figure 6.3.4 shows TEM images for the
5% loaded poly(vinyl ester) sample, PVE/CHDS-5, at both low and high magnification.
The low magnification images provide information about the nanodispersion while high
magnification images tell whether exfoliation and/ or intercalation has been achieved.
The low magnification TEM image shown on the left in Figure 6.3.4 is consistent with
formation o f a microcomposite; darker portions perceived as metal hydroxide sheets are
separated by more than 104 A. Regularly spaced stripes running diagonally across the
image are an artifact introduced by knife-edges during sample preparation. The high
magnification image shown on the right in Figure 6.3.4 does not show clear evidence of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

135

either exfoliation or intercalation, consistent with the presence of the monolayer-CHDS


and CHN phases in PVE/CHDS-5.

100

PVE/CHDS-10
PVE/CHDS- 5
PVE/CHDS- 1
PVE

20

to
o
o

PVE/CHDS-1, PVE/CHDS-5, PVE, PVE/CHDS-10

50

150

250
350
450
Temperature 1C

550

650

Figure 6.3.5 (A) TGA curves for pure PVE (dotted and dashed), PVE/CHDS-1 (solid),
PVE/CHDS-5 (hatched), and PVE/CHDS-10 (bold). (B) DTG curves for PVE/CHDS-1
(solid), PVE/CHDS-5 (hatched), pure PVE (dotted and dashed), and PVE/CHDS-10
(bold). Derivatized mass losses are scaled by a factor of 100.

Figure 6.3.5A shows thermogravimetric analysis (TGA) curves for pure PVE and
its composites at 1, 5, and 10% loading. Figure 6.3.5B shows the derivatives of these
curves. The TGA for pure PVE, in Figure 6.3.5A, consists primarily of a single step with
a small low temperature shoulder that is more obvious in the derivative curve shown in

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

136

Figure 6.3.5B. Similar trends are observed for the 1 and 5% loaded composites, but the
temperatures of maximum degradation are slightly lower than for the pure PVE.
However, the 10% loaded sample shows a different thermal behavior with two distinct
degradation steps having maximum degradation rates occurring at 340 and 430 C
respectively. The earlier mass loss may be due to the loss of the additive and/ or
monomeric units remaining after the curing process.

Table 6.3.1

TGA data for PVE composites

Sample
PVE
PVE/CHDS-1
PVE/CHDS-3
PVE/CHDS-5
PVE/CHDS-7
PVE/CHDS-10

Char (%)
70
111 (8)
130 (9)
130 (10)
130 (11)
182 (12)

Tio (C) T50 (C) AT50 (C)


4100
4440
0
4381
3856
-61
3741
4390
-50
3851
4411
-31
3423
4411
-31
2846
4401
-41

temperature at which 10% mass loss occurs; 5 temperature at which 50% mass
loss occurs; AT50, T50(composite) minus T50(pure PVE). Italicized entries are the
expected char based on the residue obtained from pure PVE and CHDS samples.
T i o ,

o ,

The initial degradation temperatures, measured by the temperature at which 10%


weight loss occurs, Ti0, are presented in Table 6.3.1 and are useful for assessing low
temperature thermal stability of composites. Ti0 values decrease with an increase in the
amount of the additive. This is clearly seen from the mass difference curves of the
composites relative to the pure PVE; (remaining mass % of PVE composites minus
remaining mass % of pure PVE at the same temperature) shown in Figure 6.3.6A. For all
three composites considered here, a destabilization behavior is seen in the temperature
range o f 100 - 450 C (peaking at 435, 420, and 400 C for 1, 5, and 10% loadings

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

137

12
PVE/CHDS-10

PVE/CHDS- 5
PVE/CHDS- 1

w
w

CD

-12

-18
5

0.5

-0.5

PVE/CHDS-10
PVE/CHDS- 5
/ PVE/CHDS- 1
PVE

50

150

250

350

450

550

650

Temperature /C

Figure 6.3.6 (A) Curves of mass loss differences for PVE composites at PVE/CHDS-1
(solid), PVE/CHDS-5 (hatched), PVE/CHDS-10 (bold) loadings as a function of
degradation temperature. (B) DTA curves for pure PVE (dotted and dashed),
PVE/CHDS-1 (solid), PVE/CHDS-5 (hatched), and PVE/CHDS-10 (bold).

respectively). The samples loaded at 1 and 5% show small and comparable mass losses in
this temperature range while a significant destabilization effect is noted for a

10

% loaded

sample. The ease with which highly loaded PVE samples degrade at lower temperatures
suggest that the additive may have a negative effect on the cross linking density of the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

138

cured composite. The presence of CHDS was also observed to lead to an increase in the
time required for the polymer composite to set. CHDS may therefore decrease the curing
reactivity o f the vinyl resins, leading to the formation o f long polymer chains with fewer
cross-links and hence low thermal stability in the lower temperature regime.
At 450 C and above, the mass loss difference curves shown in Figure 6.3.6 A are
positive for all three composites. This suggests a stabilizing behavior due to the presence
of additives, possibly catalytic formation of cross-linked char and less volatile products at
these high temperatures. The temperature at which 50% mass loss occurs, T5 0 , are slightly
lower (small negative AT50 values) for the composites compared to the virgin polymer as
shown in Table 6.3.1. These reductions in the T50 values are relatively small in
comparison to phosphorus containing PVE nanocomposites.219
The char % as measured by the amount of residue remaining at 650 C increased
with percent additive. Also shown in Table 6.3.1, are the expected char % if the residue
were additive, based on the residue obtained from pure PVE and CHDS. The fact that the
observed residue is higher than the calculated demonstrates the effectiveness of the
additive, CHDS, in char formation. Taking into account the thermal stability at low
temperatures and also the amount of char remaining at the end of the combustion process,
the 5% loaded composite provides the best stabilization of the formulations tested here.
Figure 6.3.6B shows differential thermogravimetric analysis (DTA) curves for
pure PVE and its 1, 5, and 10%-loaded composites. The DTA profile o f the pure polymer
shows a pronounced endothermic step at 450 C corresponding to its maximum
degradation rate temperature. The DTA curves for the composites show significantly
different patterns. Pronounced exotherms are seen in the region o f 250 - 420 C for the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

139

composites and may be attributed to the thermo-oxidation of the dodecyl sulfate. The
endothermic process at around 450 C is decreased in magnitude with an increase in the
percent additive. This suggest that even though there is significant mass loss at lower
temperatures for the composites, there is a compensatory stabilization process at 450 C
which happens to be the maximum degradation rate temperature for pure PVE. The
endothermic peak at around 450 C is shifted to slightly lower temperatures for the
composites with an increase in percent additive consistent with the slight reduction in
degradation temperatures as noted earlier.

1000

PVE/CHDS-5

CM

650

; PVE

CH

300
PVE/CHDS-1
PVE/CHDS-10

-50
0

50

100

150

200

250

time Is

Figure 6.3.7 Heat release rate curves for PVE (dotted and dashed), PVE/CHDS-1 (solid),
PVE/CHDS-5 (hatched), and PVE/CHDS-10 (bold) from cone calorimetry at 35 kW/m2.

The heat release rate (HRR) as a function of time for pure PVE and its composites
(1, 5, and 10% loadings) at a flux of 35 kW/m2 are shown in Figure 6.3.7. The addition of
CHDS lowers the time to sustained combustion o f the composites relative to the virgin
polymer but the HRR curves for the composites show that the evolution of heat is spread

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

over a narrow range of the combustion time. Even though the polymer composites start to
bum earlier, they extinguish much faster than the pristine PVE. There is very little or no
reduction in PHRR for all the modified PVE composites shown in Figure 6.3.7. However,
PHRR percent reductions of 18 12 and 29 6 were observed for the 3 and 7% loaded
composites. There is no obvious correlation between the reduction in PHRR and the
percent additive. Table 6.3.2 gives a summary of the results obtained for PVE and its
composites with different additive loadings.
Significant percent reductions, in the 20-30% range, in the total heat released are
observed with these low additive loading percentages. This is not uncommon, as PVE
composites containing phosphoms-based fire retardants have been shown to have lower
THRs but a wider heat release distribution profile.219 Inorganic/organic modifiers bum
out at low temperature, producing inorganic metal oxide residues that might slow or
inhibit thermal degradation of the composites. The presence of the inorganic/organic
modifiers in the PVE composites may provide heat sinks during combustion, perhaps via
the formation o f a physical barrier (cross-linked char) that prevents diffusion o f radicals
and improves the thermal stability o f the polymer.
Table 6.3.2

Cone calorimetry data for PVE composites at 35 kW/m2.

TSC PHRR (kW/m2)


(%red)
(s)
PVE
659
895+67
PVE/CHDS-1
49+5
953+49 (-7)
PVE/CHDS-3 40+2
731+80(18)
PVE/CHDS-5
286
879+56 (6)
PVE/CHDS-7 26+2
631+57(29)
PVE/CHDS-10 26+3
853+18 (5)
Sample

THR (MJ/m2) AMLR CY


(%red)
(g/stm2) (%)
137+7
27+1
4
79+1
121+10
63+1 (19)
30+1
5
101+14
62+1 (21)
25+3
6
64+1 (18)
106+8
28+1
8
62+1 (21)
34+1
101+8
9
56+3
58+1 (27)
27+2
12
T pH R R
(S )

ASEA
(m2/kg)
1089+88
1034+15
1107+41
1021+40
858+23
948+21

TSC, time to sustained combustion; PHRR, peak heat release rate (% red); TPHrr, time to
peak heat release rate; THR, total heat release (% red); AMLR, average mass loss rate;
CY, char %; ASEA, average specific extinction area (a measure of smoke).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

141

Expected percent reductions in total heat release (THR) were calculated based on
cumulative THR values assuming there is no interaction between the polymer and the
additives. Figure 6.3.8 shows plots of the expected reductions in THR, shown as open
circles, and the experimentally observed values shown as solid circles. The
experimentally obtained reductions in THR are significantly larger than the calculated
values suggesting that THR reductions are not simply a result o f replacing some fraction
o f PVE (THR = 79 MJ/m2) with an equal amount o f the additive, CHDS, which has a
lower THR value of 51 MJ/m2 for the same mass (30 g) sample. This is clear evidence
that the addition of the layered hydroxy salt, CHDS, is effective in reducing the
flammability of the PVE. The improvement in fire retardancy is shown by the reduction
in the THR and the increase in the char remaining after complete combustion as shown in
Table 6.3.2.

30
DC
X
c
o
o
TD
CD
DC

20

10

0
0

4
6
8
% CHDS Additive

10

Figure 6.3.8 Percent reduction in total heat release (THR) vs. % additive; observed
reductions () and calculated reductions (o). Calculated values assume no interaction
between PVE and CHDS.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

142

32000

Cone residue (x 0.5)

24000

650

>>

450

C
16000
0

400

350

8000

300
250

10

20

30

40

50

60

70

2 theta (degrees)

Figure 6.3.9 XRD pattern of residue after heating PVE/CHDS-10 to indicated


temperatures (250, 300, 350, 400, 450, and 650 C) at 20 C/min in TGA and residue
from cone calorimetry (top trace) experiments. The XRD pattern for the PVE/CHDS-10
cone residue reveals the presence of CuO (0), CU2 O (A), metallic Cu () reflections. An
unidentified phase (+) is also observed in the TGA residue trace at 350 C. Data is offset
for clarity.

Differences in copper oxidation states in cone calorimetric residues were observed


for PMMA composites made with CHM and a zinc/copper methacrylate HDS.52 In order
to investigate the evolution of copper oxidation states, the composite, PVE/CHDS-10 was
heated in the TGA at 20 C/min and in the cone calorimeter and the resulting residues
were analyzed by XRD; patterns are shown in Figure 6.3.9. The XRD pattern of
PVE/CHDS-10 from TGA at 250 C reveals the presence of the CHN phase and CU2 O
(PDF# 35-1091).147 The monolayer CHDS phase is not evident from XRD patterns at this
temperature, suggesting its collapse and/ or existence in an amorphous state. CU2 O and
metallic copper (PDF# 4-836)147 are seen in the XRD pattern o f the residue collected at
300 C. Of particular interest is the disappearance of the Cu (I) oxide at 350 C with the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

143

formation o f metallic copper and a second phase marked with crosses that could not be
identified. Disproportionation of Cu (I) may lead to the formation of metallic copper, Cu
(0) and Cu (II) as illustrated by the equation; 2 Cu (I) Cu (0) + Cu (II). This suggests
that the unidentified phase may be a Cu (II) containing compound, however further work
is necessary to assign this phase.
Metallic copper (Cu), generated from copper (I) salts, (CuCl, CuBr, and Cul) and
Cu (II) complexes during pyrolysis, has been reported156157to have a stabilizing effect on
the thermal degradation of poly(vinyl chloride) (PVC), by catalyzing intermolecular
cross-linking of conjugated polyenes produced during the initial stages o f degradation.
Otherwise, these polyenes could undergo cyclization reactions leading to the formation of
benzene and other aromatics, which would bum to produce heat and smoke. It is
proposed that a similar reductive coupling reaction occurs, following the abstraction of
pendant hydroxy groups from the PVE chains by metallic copper. The presence of
metallic copper in the XRD pattern o f TGA residues for PVE/CHDS-10 at 300, 350, 400,
and 450 C may prevent depolymerization of PVE through reductive coupling, thus
promoting char formation. Cu (II) readily reduces to Cu (I) or Cu(O);220 thus the
stabilization effect observed stems from the ability of Cu (II) to form zero- or low- valent
metal species upon pyrolysis. More work is required to elucidate the role o f metallic
species in the stabilization o f the polymer.
The XRD patterns o f the TGA residue at the indicated temperatures for
PVE/CHDS-10 do not show any evidence o f polycrystalline Cu(OH)2 , the expected
product of hydroxyl abstraction from polymer chains. The absence of Cu(OH ) 2

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

144

reflections in the XRD patterns of the TGA residue may suggest that it exists in an
amorphous phase.

650

0.3

450
400
350

<D 0.2
O
C

300

CO
_Q

250

O
cn

PVE/CHDS-10

_Q

<

PVE

4000

3500

3000 2500 2000 1500


Wavenumber (cm'1)

1000

500

Figure 6.3.10 Fourier transform infrared (FTIR) traces for pure PVE, PVE/CHDS-10, and
TGA residue o f PVE/CHDS-10 heated to different temperatures indicated in the plot.
Data are offset for clarity but otherwise not scaled.

Figure 6.3.10 shows the Fourier transform infrared (FTIR) of the TGA residue of
PVE/CHDS-10 at various temperatures. A sharp, intense absorption band observed
around 3740 cm'1 (for traces at 350 and 400 C) is consistent with the presence of non
hydrogen bonded Cu-OH groups.40 There is no evidence of an increase in the C = C
absorption band at around 1600 cm'1 (between 350 and 450 C) suggesting that reductive
coupling prevails over dehydration. Pike and coworkers157 reported that low-valent metal
additives prevent cracking of hydrocarbon char at high temperature, suggesting that less
volatile molecules are produced reducing smoke while increasing the remaining char.
This is consistent with reduction in the amount of smoke as seen from cone calorimetry.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

The presence of Cu ions in the polymeric matrix during combustion may serve to
stabilize the composite materials, consistent with the enhanced residue from TGA and
cone. Most likely these LHSs act as flame-retardants through three modes of action: (a)
dilution of combustion atmosphere by non-combustible gases (H 2 O and CO2 ) generated
from the thermal degradation o f the additives; (b) the inorganic combustion products may
act as a physical barrier (cross-linked char) reducing diffusion of combustible matter
(mass transport) and1 or energy transfer, hence improving flame retardation; and (c) the
metal species participate in coupling reactions preventing the generation of small
molecules that act as fuels.

6.4

Conclusions
Copper hydroxy dodecyl sulfate (CHDS) has been used as an additive to provide

thermal stability to PVE. In TGA experiments, only minor decrease in T5o is observed
with this additive, in contrast to the more significant destabilization observed with
phosphorus containing additives.221 Addition of small amounts of the CHDS (1 - 10%) to
pure PVE result in a significant reduction in the total heat release (THR) from cone
calorimetry (20 - 30%) and ~ 160% increase in the amount o f char remaining from TGA
experiments at 650 C. Similar increments (~ 200%) in the char remaining after
combustion in a cone calorimetry are observed, suggesting significant retardation in
flammability of the composites relative to pure PVE. However, there is no reduction in
peak heat release rate (PHRR). The presence of metallic Cu during the course of
degradation o f the composites is likely to catalyze the formation of cross-linked char
from reductive coupling reactions, resulting in effective thermal stabilization o f these

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

146

polymeric materials. Copper-based additives are thus very promising for the
improvement of thermal stability and fire retardancy of PVE. Studies o f possible
synergistic effects between CHDS and conventional fire retardants will be the focus of a
separate publication.

771

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

147

Chapter 7

7.1

Summary and Future Work

Summary
The work that has been carried out in our laboratory over the past 40 months has

involved several interesting and significant areas including thermodynamics and kinetics
of anionic exchanges, anion release profiles, thermal degradation o f layered 2-D
nanostructured materials, and preparation and subsequent physical property
characterization o f polymer/inorganic-organic composites. Effort has been and will
continue to be directed towards understanding the chemical structure of 2-D
nanostructured materials, HDSs and LHSs, in order to gain fundamental insights into
their reactivities. The long term goal is to optimize these materials for selected
applications in areas such as fire retardancy, magnetism, and drug delivery. In the initial
stages of the work described in chapters 2 and 3 we focused almost exclusively on
understanding the chemical reactivities as a function of structural morphology for the
model layered compounds.
A hydroxy double salt, zinc nickel acetate, was found to have a structure similar
to that of a layered hydroxy salt, zinc hydroxy acetate, where acetate anions have been
predicted from FTIR analysis to have unidentate coordination to the Zn2+ ions. The
interlayer spacing for ZNA and ZHA are comparable. Zinc copper hydroxy acetate is
structurally different from ZHA and ZNA with acetate anions bound in a bridging
fashion. The interlayer spacing of ZCA is significantly lower than that for ZHA and ZNA
suggesting possible variations in reactivities for these model compounds. The relative
ease with which acetate anions were displaced from ZHA and ZNA in comparison to

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

148

ZCA clearly shows that the reactivities of the compounds are directly related to their
initial morphology. While most of the anionic exchange reactions followed a deceleratory
temporary release profile, formate in ZCA gave a linear variation of acetate with time.
This is the first time were such a behavior has been observed with layered compounds of
this nature.
The thermal degradation o f the model compounds has been investigated. In all
cases water is lost in the lower temperature regime and this is followed by
dehydroxylation and deanation processes. Thermal degradation o f ZHA and ZNA
released acetic acid, CO2 , and significant amounts of acetone. Formation o f acetone in the
gas phase has been attributed to the ketonization of acetic acid. While acetone was
evolved from the thermal decomposition of ZHA and ZNA, little or no amounts were
observed for ZCA. The presence o f ZnO in the products of thermal degradation of ZHA
and ZNA has been implicated as having a catalytic effect promoting ketonization. Even
though the residue from thermal degradation o f ZCA contains ZnO, the relatively small
amount of Zn present limits the ketonization channel. The other products from the
decomposition process such as Cu and CuO may also have a detrimental effect on the
ketonization process.
Having gathered fundamental reactivity and thermal decomposition of these
layered materials, our attention was shifted towards understanding and effectively
enhancing the role of these materials as potential fire retardants. Tailor made layered
compounds were added to selected polymer matrices at loadings varying from 2-10%.
Even at loadings as low as 2% we observed significant reductions in the total heat release
for copper containing additives in PMMA. Natural clays do not typically show a

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

149

reduction in the total heat release, however, they result in significant reductions in peak
heat release rates. On the other hand anionic clays studied in this work did not show any
significant reduction in PHRR. The formation o f nanocomposites was found not to be
necessary for improvements in some fire properties as measured by cone calorimetry.
Reductions in total heat release were observed for PMMA and PVE composites. This
suggests that the presence of 2-D layered additives in the polymer matrix prevents total
depolymerization.
Several mechanisms have been proposed to explain the improved fire properties
for polymers containing the purported fire retardants. A physical barrier effect has been
proposed since the inception of idea of making polymer nanocomposites to improve their
thermal stability. This mechanism is dependent on the formation of a ceramic barrier
which slows or prevents the rapid diffusion of oxygen and volatiles. Reactive radicals
that might have proliferated fire growth are trapped in the solid char network and this
effectively reduces the spread o f the fire. This physical barrier can also reduce heat
transfer hence protecting the polymer underneath. Radical entrapment is not restricted to
the condensed phase, as free radicals may react with the products from additive
decomposition. Also non combustible products from the decomposition of the additives
such as water, CO2 , and CO may serve to deplete O 2 supply effectively reducing the
polymer degradation. Also catalytic effects due to Cu are likely to be more important in
reducing the flammability of these polymer composites.
Thermal degradation kinetics o f polymer composites were also explored using a
multiple heating rates kinetic analysis, the Flynn-Wall-Ozawa (FWO) method. The
calculated apparent activation energies calculated were higher for the polymer

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

150

composites relative to the virgin polymer suggesting that the composites containing the
layered additives are more thermal stable. These results are consistent with the observed
positive shift in the thermal degradation profiles using TGA. Even though these kinetic
data are derived form mass loss and give no physical meaning in terms o f bond breaking
and/ or formation, they still afford us a way o f directly comparing the thermal stability o f
the composites relative to pure polymeric material.
We have been able to investigate key structural features of the layered hydroxide
salts and ultimately suggest designs for selected applications.

7.2

Proposed Future Experiments

7.2.1

Anionic Exchange Kinetics

Based on our preliminary results, we propose to address a series of objectives


designed to provide data necessary for fundamental characterization o f anion exchange
processes. These experiments will initially be carried out with Zn/Cu HDS precursors.
Three specific experimental sequences for this objective are described below.

1A.Investigate whether the unique zero-order kinetics release o f acetate in the


exchange with formate is a general characteristic of reactions where a larger
anion is exchanged with a smaller anion.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Zn/Cu-octanoate HDS will be prepared in a complete exchange process using the


Zn/Cu-acetate HDS precursor synthesized as described previously. The octanoate analog
is convenient for these studies because of the high exchange efficiency (i.e.
thermodynamic stability) of the longer alkyl chain HDS, a trend that has also been noted
in other work.30 This provides us with a base material well suited for testing the
feasibility of designing tunable release rate protocols via systematic variations in
exchange anion size. These first series o f experiments will involve exchange with
smaller n-alkyl anions (e.g., formate, acetate, propionate, butyrate) combining NMR and
PXRD analysis. PXRD data will be evaluated to determine whether the exchange
proceeds via a simple topotactic transition from starting material to a single product
phase or whether the mechanism is more complex, involving intermediate structures or
producing multiple product phases.

IB. Explore a wider range of anion size, structure, and chemical composition
effects on the observed exchange kinetics.

Investigations of the anion substitution kinetics for three classes of anion


structural alterations are proposed: (a) branched alkyl carboxylates, (b) carboxylates that
include other functional groups, and (c) other binding groups such as sulfonates. These
modifications are designed to better characterize the contributions of diffusion and
nucleation to the observed kinetics, in cases where Avarmi- Erofe'ev kinetics are
followed, as well as to lay the ground work for optimizing these materials for other
applications.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

152

While the value o f m obtained form the Avrami-Erofe'ev model is used to


distinguish between diffusion- and nucleation-control o f the exchange kinetics, it is
important to note that there are limitations inherent in making such a global interpretation
in complex systems where defects and phase boundaries may play important roles.
Systematic alterations in anion structure can, however, be combined with control
experiments, such as running reactions with the role of precursor and exchange anion
reversed and repeating experiments with separately synthesized batches o f precursor, to
obtain data which can be used to test the main assumptions o f this model.
Noting that we have observed two sets of d-spacings for butyrate and propionate,
and that the butyrate exchange kinetics may involve an intermediate solid phase structure,
we plan to perform more exchange experiments with acetate precursors and these species.
The /-butyrate anion is chosen as one of the branched alkyl carboxylates for comparison
with these other small anions. 2-ethyl hexanoate exchange kinetics will be compared
with that of w-octanoate and w-hexanoatc. Recalling that n-octanoate exchanged for
acetate exhibited simple Avarmi-Erofe'ev behavior, the results of analogous experiments
with 2-ethyl hexanoate will be examined to determine whether there is a significant
decrease in m, as would be expected for increasing contribution from diffuse effects.
Other more highly branched alkyl chains will be tested if these first experiments indicate
a trend that needs to be further characterized. In addition, we will examine whether the
number of equivalents (per HDS unit) of alkyl groups changes with use of more bulky
groups.
Variations in functional groups can be utilized to examine the effects of changing
polarity and hydrogen bonding capability on the exchange reactivity, factors that will be

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

153

important in long-term goals o f our research program to develop these materials for
selected applications such as chemical sensing and polymer fire retardancy. Lactate
reactivity can be compared with /'-butyrate. Acrylate and methacrylate can be compared
with corresponding alkyls. The proposed series of benzoate derivatives (benzoate, phydroxy benzoate, and salicylate) combines phenyl and hydroxyl groups.

co2HjC

H CHj

co2-

H2

H C^S'
3

H,

isobutyrate

Cf02

'C '' hT c ^ c h 3
H,

H,

2-ethyl hexanoate

H ,C

H OH

lactate

C 02
SO,
3
CH,

m ethane sulfonate

cpo;

CH

II

//
H2C

CH,

acrylate

CH,
3

methacrylate

>c- v O
n
benzoate

OH

salicylate

naproxen

hydroxycinnamate

p-hydroxy
b enzoate

Figure 7.2.1 1 Exchange anions.

l.C Determine temperature and exchange solution concentration effects on


observed kinetics.

In cases where the observed kinetics can be fit to a single model such as shown in
our work, temperature dependence studies can be used to extract activation energies. We
have also observed that exchange of Zn/Cu-acetate HDS with butyrate proceeds to a
single product phase when a complete exchange experiment is carried out at 40 C, in
contrast to the room temperature. We propose additional studies to characterize the
temperature dependence of product phases produced from butyrate and propionate, as

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

154

well as carry out detailed studies to examine whether the intermediate phase observed in
the butyrate exchange kinetics studies is present at other temperatures.

7.2.2

In-siiu Anion Release Profile using Ultraviolet-Visible (UV-VIS)


Spectroscopy and Synthesis of Thin Layered Films

The second goal for future work involves designing experiments that would
enable us to follow the exchange reactions in-situ. It is well known that conjugated
organic molecules absorb strongly in UV/VIS region of the electromagnetic spectrum. By
selecting a wavelength (strongest non overlapping absorption) at which a particular
organic anion can be monitored, one can easily follow the release profile of the target
anion.
In our laboratory we have designed a simple in-situ exchange kinetics experiment
to monitor the release profile o f naproxen. Naproxen is a non steroidal anti-inflammatory
drug (NSAID) used for the treatment o f arthritis and cardiovascular disorders. It would
therefore be interesting to study the release profile o f naproxen from layered materials.
Khan and coworkers60 have shown that pharmaceutically active compounds can be
reversibly intercalated into layered double hydroxide and the studies suggest application
o f these materials as basis for tunable drug delivery systems. Ambrogi and coworkers222
also studied the uptake o f the diclofenac into Mg-Al- hydrotalcite for controlled release
formulation. We note here however, that these studies are merely for extracting the
exchange reaction kinetic parameters and not for medical purposes.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

155

7.2.3

Polymer Modification
Results obtained in our laboratory have shown that these layered

inorganic/organic hybrids can be used as additives to polymers as a way of improving the


thermal stability of the polymers. A remarkable reduction in the total heat release (THR)
after addition o f these anionic clays has been noted. However, dispersion o f the layered
materials in the polymer matrix is still poor. We will continue to explore ways of
improving the additive dispersion in the polymer matrix. These materials show different
behavior as compared to natural clays and they promise to hold a key to improving fire
retardancy.
Adding the layered materials to the polymer increased thermal stability o f the
later but it is still unclear how this is achieved. By monitoring the profiles o f combustion
products of the polymer composites and analyzing the inorganic residue as we bum them,
will afford us fundamental insights into the thermal degradation mechanism of these
materials. Conventional additives will be combined with anionic clays to explore possible
synergism in polymer thermal stability or fire retardancy. There are several expectations
for the future of polymer composites which include;
(i)

Development of a clear understanding o f the role of these 2-D nanostructured


materials in improved fire retardancy,

(ii)

Mechanistic elucidation of the role of additive in improved thermal stability


and fire safety,

(iii)

Understanding why polymer composites ignite faster than virgin polymers,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

156

(iv)

Probing the influence o f extent of nanodispersion on the peak heat release rate
and total heat release.

7.2.4

Predictive Mathematical Modeling of Thermal Degradation of Polymers


In Chapter 5, we demonstrated the use o f the Flynn-Wall-Ozawa isoconversional

method to estimate the apparent activation energy as a function of conversion fraction for
the thermal degradation of polymeric material. Results obtained gave an insight into the
relative thermal stability of composites when compared to the virgin polymer. On-going
research and future work in our laboratory will focus on using geometric modeling
equations in order to be able to predict the life time or degradation profiles o f certain
composites; a very useful tool in assessment o f polymer products for their end use and
fire safety. Preliminary results obtained in our laboratory and future directions are
discussed herein.
The study of kinetics of thermal degradation o f polymers has recently emerged as
an area of interest from which thermal stability (lifetime) and mechanisms of the
decomposition process can be predicted.

'

This kind of study has been achieved in

many cases by use of TGA apparatus where the rate of mass loss is measured. An
inherent disadvantage of this method in obtaining kinetic data is that there is no direct
correlation between broken and/ or formed bonds and the rate of mass loss. However, the
overall chemical and physical processes are expected to influence mass loss rates and the
obtained kinetic parameters from mathematical modeling can be used to simulate thermal
degradation profiles. Attempts will be made to find the best modelistic descriptor of the
experimental data and use the extracted kinetic parameters to reconstruct the thermal

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

157

degradation profile. This is useful in predicting the thermal behavior of the same
polymeric system but under slightly different heating conditions.
Predictive modeling has been tested with poly (ethylene) (PE) polymer containing
5% mass fraction of a layered hydroxy double salt, zinc copper hydroxy stearate (ZCHS).
This composite is identified as PE-ZCHS-5. The additive, ZCHS, was prepared by anion
exchange were ZCA HDS (5 g) was mixed with 0.1 M sodium stearate solution in
chloroform (500 mL) for 24 h at room temperature with frequent shaking. The
supernatant was decanted and replaced by a fresh solution for another 24 h with this
process repeated three times. The exchanged HDS product, zinc copper hydroxy stearate
was recovered by filtration, washed with 500 mL o f methanol, followed by 500 mL o f
acetone and plentiful amounts o f water before drying at 100 C overnight to drive off
undesirable solvents especially chloroform.
PE-ZCHS-5 was prepared via melt blending using established methods.187 ZCHS
(5 g) was mixed with pristine poly (ethylene) to achieve a 5% mass fraction. Melt
blending was then performed on a Brabender mixer for 10 minutes at a temperature of
about 130 C and speed of 60 rpm. A reference sample o f pure poly (ethylene) was
obtained by following the same procedure only without an additive. TGA was done as
discussed in Chapter 6 except that a heating rate of 10 C/min was employed.
Thermogravimetric analysis (TGA) decomposition behaviors of pure poly
(ethylene) and PE-ZCHS-5 in air are shown in Figure 7.2.4.1. Corresponding TGA data
are presented in Table 7.2.4.1. The onset o f thermal degradation, measured as the
temperature at which 10% mass loss occurs, Tio, is shifted to a higher value by 59 C for
PE-ZCHS-5 relative to pure PE. The presence o f the additive, ZCHS, delays the initiation

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

158

100

Pure PE

PE-ZCHS-5%

m 60
CO

cc
^

50

150

250

350

450

550

Temperature /C
Figure 7.4.2.1 TGA curves for pure PE (empty circles) and PE-ZCHS-5 (solid line).

Table 7.4.2.1 TGA data for ZCHS, poly(ethylene) and PE-ZCHS-5


Sample
Pure PE
PE-ZCHS-5
ZCHS

Tio (C)
3604
4193
3181

T50(C) a t 50 (C) Char %


421 3
0
00
4464
255
10
4690
25 1
-

Tio, temperature at which 10% mass loss occurs; T5o, temperature at which 50% mass loss occurs;
AT50, T50(composite) minus T50(pure PE).

o f thermal degradation. This may be due to the physical barrier presented by the lamellar
platelets between which polymer units may be sandwiched. Temperature shifts to higher
magnitudes for the composite relative to the pure PE are seen throughout the degradation
process. This is clearly illustrated in the mass difference curve of PE-ZCHS-5 relative to
that of pure PE; (remaining mass % of PE-ZCHS-5 minus remaining mass % o f pure PE
at the same temperature) shown in Figure 7.2.4.2. PE-ZCHS-5 is significantly stabilized

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

159

in the temperature region of 325 - 600 C with the maximum stabilization achieved at
about 410 C.

PE-ZCHS-5%

CO
CO

cu
2
<

50

150

250

350

450

550

Temperature /C
Figure 7.4.2.2 Mass loss difference curve for PE-ZCHS-5 (solid line) as a function of
degradation temperature.

T 5o,

the temperature at which 50% mass loss occurs, is shifted to a higher value

by 25 C for PE-ZCHS-5 relative to pure PE. No meaningful destabilization is seen below


325 C to suggest that the presence of ZCHS at these loading percentages has very little
negative effect in the low temperature regime. This is in contrary to observations
previously made in our laboratory when PVE was stabilized with copper hydroxy
dodecyl sulfate (CHDS) at mass fractions of 10% or less.221 Stearate groups have a
similar chemical structure to the PE chains and are thus expected to have little if any
unfavorable effects on the thermal stability o f PE. Formation o f char may serve as a
physical/chemical barrier preventing or slowing down the diffusion of volatile gases,
which would otherwise act as fuels resulting in the rapid propagation of the degradation

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

160

processes. The anticipated barrier also prevents diffusion of oxygen to the polymer
matrix hence retarding the depolymerization process.
Thermal degradation of pure PE and its composite can roughly be envisioned as a
single step process. Dynamic analysis of the thermal degradation process provide means
to elucidate mechanism o f solid state reactions and to extract reliable and precise
estimates of kinetic parameters such as activation energies, frequency factor, rate o f
decomposition, and in some cases the order o f the decomposition reaction. Even though
model fitting methods have been extensively criticized for non-isothermal applications
due to resultant indistinguishable fits between different models,195196their importance
can not be overestimated as they may provide information on energy barriers and offer
mechanistic clues. One can directly calculate the kinetic parameters from a single nonisothermal run something that cannot be achieved via isoconversional methods. The
calculated kinetic parameters are then used to construct decomposition simulation data to
be compared against experimental data. Thus the parameters obtained are useful in the
prediction of thermal behavior for PE and its composites.
The basic equation for the thermal decomposition assuming a solid-state reaction
is described by,
da = k fra)
dt

7.4.2.1

where k is the reaction rate constant, f(a) the reaction model, and a is the conversion
fraction. The rate constant, k, is related to the reaction temperature by the Arrhenius
equation,
k = Ae'Ea/RT

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

7.4.2.2

161

where A is the pre-exponential factor (min'1), Ea is the apparent activation energy (kJ
mol'1), and R is the gas constant (J/ molK). Substituting Eq. 7.4.22 into Eq. 7.4.2.1
yields,
da = Ae'Ea/RT f(a)
dt

7.4.2.3

The use of Eq. 7.4.2.3 assumes that the triplet (A, Ea, and f(a)) describes physical and/ or
chemical changes with accuracy, during the thermal degradation of polymeric material Given
that,
da = da.dt
dT dt dT

7.4.2.4

and dT/dt is the heating rate, P (10 K/min), substituting Eq. 7.4.2.4 into Eq. 7.4.2.3 will
yield the differential equation below,
da = A e'Ea/RTf(a)
dT P

7.4.2.5

On integration, Eq. 7.4.2.5 gives,

g(a) = A Jo e'Ea/RT dT

7.4.2.6

P
Replacing Ea/RT by x and transforming the limits of integration can further simplify
Eq. 7.4.2.6 above to yield Eq. 1.4.2.1 below,
00
g(a) = A E a J, d jc
pR x2

7.4.2.7

which can be further written as,


g(a) = AEae;x h(x)
PR x2

where h(x) is the exponential integral represented by,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

7.4.2.8

162

h(x) =

x4 + 18x3 + 86x2 + 96x


x4 + 20x3 + 120x2 + 240 x + 120

which is a Senum and Yang approximation.226 In a case where the kinetic model is
known in priori, Eq. 7.4.2.9 can be applied to calculate the activation energy and
frequency factor.
ln(g(a)/T2) = [ln(AR/Ea) - In/?] - Ea/RT

7.4.2.9

Several models corresponding to different physical and chemical processes


(Appendix A) were fitted to the experimental data using linear regression analysis. The
kinetic model, which gave the highest correlation coefficient, 3-D diffusion-Jander
equation, was chosen as the best descriptor of the degradation mechanism. The chosen
method is used here solely for the determination and comparison of kinetic parameters
hence correlation to thermal stability rather than for the purpose o f finding a universally
perfect fit to the experimental data. The integral form o f the equation is shown below,
g(a) = [l-(l-a )173]2

7.4.2.10

The 3-D diffusion-Jander equation produced kinetic data (A and Ea) for each
simulation. The average activation energy is calculated to be 155 1 kJ/mol while the
frequency factor is 6.4 0.1 x 109 min'1 (r2= 0.9897) for purePE. When loaded with 5%
ZCHS, the activation energy of the PE composite increased to 359 5 kJ/mol and the
frequency factor was found to be 2.8 0.0 x 1024 min'1 (r2 = 0.9889). Even though the
same kinetic model was used for both PE-ZCHS-5 and pure PE, a direct comparison of
their activation energies alone as a measure of thermal stability does not have a physical
meaning since their frequency factor values are different.
To estimate the thermal stability of the these polymeric materials, we can
calculate the compensation parameter, S*p, from Ea/log A. Pure PE has a larger S*p value

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

163

a
c

0 .8

o
o

C
O
!_
LL.

0 .6
Pure PE

0.4
CD

PE-ZCHS-5%

02

50

150

250

350

450

550

Temperature /C
Figure 7.4.2. 3 Experimental (solid lines) and reconstructed a-T curves for the nonisothermal degradation of pure PE (empty circles) and PE-ZCHS-5 (empty triangles)
respectively.
of 15.8 0.1 x 103 as compared to PE-ZCHS-5 which has a value o f 14.7 0.2 x 103.
This suggests that the 5% loaded PE is thermally more stable than the virgin polymer.
Figure 7.4.2.3 shows both the experimental and reconstructed a-T plots for pure PE and
PE-ZCHS-5. The reconstructed a-T plots show a dissent match to the experimental plots
an indication that the calculated kinetic parameters are reliable.
Model fitting affords one to reconstruct thermal degradation profiles using the
obtained kinetic parameters. We will obtain TGA profiles for PE-ZCHS-5 using slightly
different heating rates and then compare those experimental data to data obtained from
mathematical reconstruction using parameters previously calculated for the 10 C/min
ramp rate. This way we hope to be able to construct and hence predict the degradation
profile of a given composite without performing the actual experiments.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

References
1 Rives, V. In Layered Double Hydroxides: Present and Future; Nova: New York, USA,
2001.
2 Reichle, W. T. J. Catal. 1985, 94, 547.
3 Kappenstein, C.; Cemak, J.; Brahmi, R.; Duprez, D.; Chomic, J. Thermochim. Acta.
1996, 276, 65.
4 Cemak, J.; Chomic, J.; Kappenstein, C.; Brahmi, R.; Duprez, D. Thermochim. Acta.
1996, 276, 209.
4 Guo, Y.; Weiss, R.; Epple, M. Eur. J. Inorg. Chem. 2005, 3072.
5 Crepaldi, E. L.; Pavan, P. C.; Valim, J. B. Mater. Chem. 2000,10, 1337.
6 Morioka, H.; Tagaya, H.; Kadokawa, J.; Chiba, K. Recent. Res. Devel. In Mat. Sci.
1998,1, 137.
7 Bruschini, C. S.; Hudson, M. J. Progress in Ion Exchange Special Publication - Royal
Society o f Chemistry, 1997,196, 403.
8 Meyn, M.; Beneke, K.; Lagaly, G. Inorg. Chem. 1993, 32, 1209.
9 Su, S.; Wilkie, C.A. Poly. Chem. 2003, 41, 1124.
11 Tyan, H.; Leu, C.; Wei, K. Chem. Mater. 2001,13, 222.
12 Wang, D.; Zhu, J.; Yao, Q.; Wilkie, C. A. Chem. Mater. 2002,14, 3837.
13 Zhu, J.; Morgan, A. B.; Lamelas, F. J.; Wilkie, C. A. Chem. Mater. 2001,13, 3774.
14 Chen, B.; Evans, J. R. G. J. Phys. Chem. B. 2004,108, 14986.
15 Xie, W.; Gao, Z.; Pan, W.; Hunter, D.; Singh, A.; Vaia, R. Chem. Mater. 2001,13,
2979.
16 Berta, M.; Lindsay, C.; Pans, G.; Camino, G. Poly. Degrad. Stab. 2006, 91, 1179.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

165

17 Zhang, S.; Horrocks, R. A.; Hull, R.; Kandola, B. K. Poly. Degrad. Stab. 2006, 91,
719.
18 Zheng, X.; Jiang, D. D.; Wang, D.; Wilkie, C. A. Poly. Degrad. Stab. 2006, 91, 289.
19 Zhang, J.; Jiang, D. D.; Wilkie, C. A. Poly. Degrad. Stab. 2006, 91, 298.
20 Fujita, W.; Awaga, K. Inorg. Chem. 1996, 35, 1915.
21 Laget, V.; Homick, C.; Rabu, P.; Drillon, M. Mater. Chem. 1999, 9, 169.
22 Fujita, W.; Awaga, K. J. Am. Chem. Soc. 1997,119, 4563.
23 Kwon, T.; Tsigdinos, G. A.; Pinnavaia, T. J. J. Am. Chem. Soc. 1988,110, 3653.
24 Schollhom, R. In: Atwood, J. L.; Davies, J. E. D.; MacNicol, D.D. (Editors), Inclusion
Compunds. 1984, p245.
25 Eldeman, C. H.; Favejee, J. C. L. ZeitKryst, 1940,102, 417.
26 Rajamathi, M.; Thomas, G. S.; Kamath, P. V. Proc. Indian Acad. Sci. (Chem Sci.),
2001,775, 671.
27 Yeh, J.; Liou, S.; Lai, C.; Wu. P.; Tsai, T. Chem. Mater. 2001,13, 1131.
28 Constantino, V. R. L.; Pinnavaia, T. J. Inorg. Chem. 1995, 34, 883.
29 Hu, C.; Li, D.; Guo, Y.; Wang, E. Chinese Sci. Bull. 2001, 46, 1061.
30 Alberti, G.; Costantino, U. Layered Solids and Their Intercalation Chemistry. Alberti,
G., Bein, T., Eds.; In Comprehensive Supramolecular Chemistry: Two and Three
Dimensional Networks', Elsevier Science Ltd: Great Britain, 1996.
31 Leroux, F.; Besse, J. Chem. Mater. 2001,13, 3507.
32 Bubniak, G. A.; Schreiner, W. H.; Mattoso, N.; Wypych, F. Langmuir, 2002,18, 5967.
33 Chen, B.; Qu, B. Chem. Mater. 2003,15, 3208.
34 Chei, W.; Feng, L.; Qu, B. Chem. Mater. 2004,16, 368.
35 Cavani, F.; Trifiro, F.; Vaccari, A. Catal. Today, 1991, 77, 173.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

166

36 Nishizawa H.; Yuasa, K. J. Solid State Chem. 1998,141,229.


37 Rojas, R.; Barriga, C.; Ulibarri, M. A.; Malet, P.; Rives, V. J. Mater. Chem. 2 002,12,
1071.
38 Stahlin, W.; Oswald, H. R. Acta Cryst. B. 1970, 26, 860.
Morioka, H.; Tagaya, H.; Karasu, M.; Kadokawa, J.; Chiba, K. Inorg. Chem. 1999, 38,
4211.
40 Masciocchi, N.; Corradi, E.; Sironi, A.; Moretti, G.; Minelli, G.; Porta, P. J. Sol. State
Chem. 1997,131, 252.
41 Jim'enez-L'opez, A.; Rodr'iguez-Castell'on, P.; Olivera-Pastor, P.; Maireles-Torres, P.;
Tomlison, A. A. G.; Jones, D. J.; Rozie're, J. Mater. Chem. 1993, 3, 303.
42 Yamanaka, S.; Sako, T.; Hattori, M. Chem. Lett. 1989, 1869.
43 Yamanaka, S.; Sako, T.; Seki, K.; Hattori, M. Solid State Ionics. 1992, 53-56, 527.
44 Bovio, B.; Locchi, S. J. Cryst. Spec. Res. 1982,12, 507.
45 Lee, S.; Her, Y.; Matijevic, E. J. Colloidal and Interface Science. 1997,186, 193.
46 Jahn, H.; Teller, E.; Proc. Roy. Soc. 1931, A161, 220.
47 Tanaka, H.; Terada, S. J. Thermal. Anal. 1993, 39, 1011.
48 Anbarasan, R.; Lee, W. D.; IM, S. S. Bull. Mater. Sci. 2005, 28, 145.
49 Tezuka, S.; Chitrakar, R.; Sonoda, A.; Ooi, K.; Tomida, T. Green Chem. 2004, 6, 104.
50 Bhaumik, A.; Samanta, S.; Mai, N. K. Indian J. Chem. Section A: Inorg, Bio-inorg,
Phys, Theor & Anal Chem. 2005, 44A, 1406.
51 Seida, Y.; Nakano, Y. Water Res. 2002, 36, 1306.
52 Messersmith, B.P.; Stupp, S.I. Chem. Mater. 1995, 7, 454.
53 Kandare, E.; Hall, D.; Jiang, D. D.; Hossenlopp, J. M. In Fire and Polymers IV:
Materials and Concepts fo r Hazard Prevention, Eds. Wilkie, C. A.; Nelson, G. L.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

ACS Symposium Series, Oxford University Press, 2006, 131.


54 Kandare, E.; Deng, H.; Wang, D.; Hossenlopp, J. M. Polym. Adv. Technol. In Press.
55 Khan, A. I.; O Hare, D. J. Mater. Chem. 2002,12, 3191.
56 Mohanambe, L.; Vasudevan, S. Inorg. Chem. 2005, 44, 2128.
57 Newman, S. P.; Jones, W. New J. Chem. 1998, 105.
58 Suzuki, N.; Nakamura, Y.; Watanabe, Y.; Kanzaki, Y. Chem. Pharm. Bull. 2001, 49,
964.
59 Mohanambe, L.; Vasudevan, S. J. Phys.Chem. B. 2005,109, 15651.
60 Khan, A. I.; Lei, L.; Norquist, A. J.; O Hare, D. Chem. Commun. 2001, 2342.
61 Alexandre, M.; Dubois, P. Mater. Sci. Eng. 2000, R28, 1.
62 Gu, A.; Liang, G. Poly. Degrad. Stab. 2003, 80, 383.
63 Zanetti, M.; Camino, G.; Canavese, D.; Morgan, A. B.; Lamelas, F. J.; Wilkie, C. A.
Chem. Mater. 2002,14, 189.
64 Choi, J.; Harcup, J.; Yee, A. F.; Zhu, Q.; Laine, R. M. J. Am. Chem. Soc. 2001,123,
11420.
65 Kommann, X.; Thomann, R.; Mulhaupt, R.; Finter, J.; Berglund, L. J Polym Sci Part
A: Polym Chem. 2002, 86, 2643.
66 Devax, E.; Rochery, M.; Bourbigot, S. Fire Mater. 2002, 26, 149.
67 Kim, D.-S.; Kim, J.-T.; Woo, W.-B. J. Appl. Poly. Sci. 2005, 96, 1641.
68 Yuan, Q.; Wei, M.; Evans, D. G.; Duan, X. J. Phys. Chem. B. 2004,108, 1238.
69 Rousselot, I.; Taviot-Gueho, C.; Leroux, F.; Leone, P.; Palvadeau, P.; Besse, J-P. J.
Solid State Chem. 2002,167, 137.
70 Velu, S.; Suzuki, K.; Hashimoto, S.; Satoh, N.; Ohashi, F.; Tomura, S. J. Mater. Chem.
2001,11, 2049.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

168

71 Prevot, V.; Casal, B.; Ruiz-Hitzky, E.; J. Mater. Chem. 2001,11, 554.
72 Rives, V.; Kannan, S. J. Mater. Chem. 2000,10, 489.
73 Aramendia, M. A.; Aviles, Y.; Borau, V.; Luque, J. M.; Marinas, J. M.; Ruiz, J. R.;
Urbano, F. J. J. Mater. Chem. 1999, 9, 1603.
74 Labajos, F. M.; Sastre, M. D.; Trujillano, R.; Rives, V. J. Mater. Chem. 1999, 9, 1033.
75 Prevot, V.; Forano, C.; Besse, J. P.; Abraham, F. Inorg. Chem. 1998, 37, 4293.
76 Barriga, C.; Jones, W.; Malet, P.; Rives, V.; Ulibarri, M. A. Inorg. Chem. 1998, 37,
1812.
77 Zhao, H.; Vance, G. F. Inorg. Chem. 1997,11, 1961.
78 Kooli, F.; Rives, V.; Ulibarri, M. A. Inorg. Chem. 1995, 34, 5114.
7Q

Kuwahara, T.; Onitsuka, O.; Tagaya, H.; Kadokawa, J.; Chiba, K. Journal o f Inclusion
Phenomena and Molecular Recognition in Chemistry. 1994,18, 59.

80 Choy, J.; Kwon, Y.; Song, S.; Chang, S. Bull. Korean Chem. Soc. 1997,18, 450.
81 O Brien, S.; Francis, R. J.; Fogg, A.; OHare, D.; Okazaki, N.; Kuroda, K. Chem.
Mater. 1999,11, 1822.
82 Fogg, A. M.; O Hare, D. Chem. Mater. 1999,11, 1771.
O'!

OHare, D.; Evans, J. S. O.; Price, S. Molecular Recognition and Inclusion, 1998, 153.

84 Williams, G. R.; Norquist, A. J.; O Hare, D. Chem. Mater. 2004,16, 975.


85 Sheridan, A. K.; Anwar, J. Chem. Mater. 1996, 8, 1042.
86 Fogg, A. M.; Dunn, J. S.; OHare, D. Chem. Mater. 1998,10, 356.
87 Evans, J. S. O.; Price, S. J.; Wong, H.; O Hare, D. J. Am. Chem. Soc. 1998,120,
10837.
88 Chang, Z.; Evans, D. G.; Duan, X.; Vial, C.; Ghanbaja, J.; Prevot, V.; de Roy, M.;

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

169

Forano, C. J. Solid State Chem. 2005,178, 2766.


89 Guo, Y.; Zhang, H.; Zhao, L.; Li, G-D.; Chen, J-S.; Xu, L. J. Solid State Chem. 2005,
178, 1830.
90 Wei, M.; Yuan, Q.; Evans, D. G.; Wang, Z.; Duan, X. J. Mater. Chem. 2005,15, 1197.
91 Del Arco, M.; Gutierrez, S.; Martin, C.; Rives, V.; Rocha, Joao. J. Solid State Chem.
2004, 777,3954.
92 You, Y.; Zhao, H.; Vance, G. F. J. Mater. Chem. 2 0 0 2 ,12, 907.
93 Li, S-P.; Xu, J-J.; Chen, H-Y. Mater. Lett. 2005, 59, 2090.
94 Bin Hussein, M. Z.; Long, C. W. Mater. Chem. Phys. 2004, 85, 427.
95 Newman, S. P.; Jones, W. J. Solid State Chem. 1 999,148, 26.
96 Bontchev, R. P.; Liu, S.; Krumhansl, J. L.; Voigt, J.; Nenoff, T. M. Chem. Mater. 2003,
15, 3669.
97 Muramatsu, H.; Suzuki, A.; Kadokawa, J.; Tagaya, H. Transactions o f Materials
Research Society o f Japan. 2001, 26, 519.
98 Meyn, M.; Beneke, K.; Lagaly G. Inorg. Chem. 1990, 29, 5201.
99 Choy, J.; Kwon, Y.; Han, K.; Song, S.; Chang, S. Mater. Lett. 1998, 34, 356.
100 Kopka, H.; Beneke, K.; Lagaly, G. J. Colloidal and Interface Sci. 1988,123, 427.
101 Rueff, J-M, Nierengarten, J-F.; Gilliot, P.; Demessence, A.; Cregut, O.; Drillon, M.;
Rabu, P. Chem. Mater. 2004,16, 2933.
102Fujita, W.; Awaga, K.; Yokoyama, T. Inorg. Chem. 1997, 36, 196.
103 Cheary R.W.; Coelho A. A. Programs XFIT and FOURYA, deposited in CCP14.
Powder Diffraction Library, Engineering and Physical Sciences Research Council,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

170

Daresbury Laboratory, Warrington, England. (http://www.ccpl4.ac.uk/tutoriaExfit95/xfit.htm), 1996.


104 Jenkins, R.; Synder, R. L. Introduction to X-ray Powder Diffractometry, Wiley: New
York, 1996.
105 ACD/Labs UVIR Program for Spectroscopic Analysis, by Advanced Chemistry
Development Inc. 90 Adelaide St. W. Suite 702, Toronto, Canada,
http://www.acdlabs.com/.
106 Rajamathi, M.; Nataraja, G. D.; Ananthamurthy, S.; Kamath, P. V. J. Mater. Chem.
2000,10, 2754.
107

LMGP-Suite Suite of Programs for the interpretation of X-ray Experiments, by Jean


Laugier and Bernard Bochu, ENSP/Laboratoire des Materiaux et du Genie Physique,
BP 46. 38042 Saint Martin d'Heres, France. http://www.inpg.fr/LMGP and
http://www.ccpl4.ac.uk/tutoriaElmgp/.

108 Poul, L.; Jouini, N.; Fievet, F. Chem. Mater. 2000,12, 3123.
109 Dou, Y. J. Chem. Educ. 1990, 67, 134.
110 Minkova, N.; Krusteva, M.; Nikolov, G. J. Molec. Structure. 1984,115, 23.
111 Nakamoto, K. Infrared and Raman Spectra of Inorganic and Coordination
Compounds, 4th ed.; Wiley: New York, 1986; 232.
112 Deacon, G. B.; Phillips, R. J. Coord. Chem. Rev. 1980, 33, 227.
113 Nakamoto, K.; Fujita, J.; Tanaka, S.; Kobayashi, M. J. Am. Chem. Soc. 1957, 79,
4904.
114 Cao, G.; Mallouk, T. E. Inorg. Chem. 1991, 30, 1434.
115 Avrami, M. J. Chem. Phys. 1940, 8, 212.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

171

116 Avrami, M. J. Chem. Phys. 1941, 9, 177.


117 Brown, W. E . ; Dollimore, D .; Galwey, A. K. Reactions in the Solid State, In
Comprehensive Chemical Kinetics, Eds. Bamford, C. H .; Tipper, C. F. H. vol 22.;
Elsevier: Amsterdam, 1980; pp. 40-113.
118 Walton, R. I.; O Hare, D. J. Phys. Chem. B. 2001,105, 91.
119 Dutta, P. K.; Puri, M. J. Phys. Chem. 1989, 93, 376.
120 Itoh, T.; Ohta, N.; Shichi, T.; Yui, T.; Takagi, K. Langmuir. 2 003,19, 9120.
121 Vucelic, M.; Moggridge, G. D.; Jones, W. J. Phys. Chem. 1995, 99, 8328.
122 Kanoh, T.; Shichi, T.; Tagaki, K. Chem. Lett. 1999, 2, 117.
123 Kandare, E.; Chigwada, G.; Wang, D.; Wilkie, C. A.; Hossenlopp, J. M. Poly. Degrad.
Stab. In Press.
124 Lei, L.; Millange, F.; Walton, R. I.; O Hare, D. J. Mater. Chem. 2000,10, 1881.
125Lei, L.; Vijayan, P. R.; OHare, D. J. Mater. Chem. 2001,11, 3276.
126 Xu, Z. P.; Braterman, P. S.; Yu, K.; Xu, H.; Wang, Y.; Brinker, C. J. Chem. Mater.
2004,16,2750.
127 Bellamy, L. J. The Infrared Spectra o f Complex Molecules; Chapman & Hall,
London, 1975.
128 Kandare, E.; Hossenlopp, J. M. Inorg. Chem. In Press.
129 Xu, Z. P.; Zeng, H. C. J. Phys. Chem. B. 2001,105, 1743.
130 Xu, Z. P.; Braterman, P. S. J. Mater. Chem. 2003,13, 268.
131 Kandare, E.; Chigwada, G.; Wang, D.; Wilkie, C. A.; Hossenlopp, J. M. Poly. Degrad.
Stab. 2006, 91, 1209.
132 Del Arco, M.; Malet, P.; Trujillano, R.; Rives, V. Chem. Mater. 1999,11, 624.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

172

133 Chellam, U.; Xu, Z. P.; Zeng, H. C. Chem. Mater. 2000,12, 650.
134 Labajos, F. M.; Rives, V.; Malet, P.; Centeno, M. A.; Ulibarri, M. A. Inorg. Chem.
1996, 35, 1154.
135 Prevot, V.; Forano, C.; Besse, J. P. Chem. Mater. 2005, 17, 6695.
136 Kovanda, F.; Rojka, T.; Dobesova, J.; Machovic, V.; Bezdicka, P.; Obalova, L.;
Jiratova, K.; Grygar, T. J. Solid State Chem. 2006,179, 812.
137 Sileo, E. E.; Jobbagy, M.; Paiva-Santos, C. O.; Regazzoni, A. E. J. Phys. Chem. B.
2005,109, 10137.
138 Lin, Y.; Adebajo, M. O.; Frost, R. L.; Kloprogge, J. T. J. Therm. Anal. Cal. 2005, 81,
83.
139 Arii, Tadashi.; Kishi, A. Thermochimica Acta. 2003, 400, 175.
140 Kawai, Akiko.; Sugahara, Y.; Park, I. Y.; Kuroda, K.; Kato, C. Ceramic Powder
Science IV, 1991, 22, 75.
141 Vithal Ghule, A.; Lo, B.; Tzing, S.; Ghule, K.; Chang, H.; Ling, Y. C. Chem.
Physics Lett. 2003, 381, 262.
142 Bera, P.; Rajamathi, M.; Hegde, M. S.; Kamath, P. V. Bull. Mater. Sci. 2000, 23,
141.
143 Bae, H. Y.; Choi, G. M. Sensors and Actuators B: Chemical. 1999, 55, 47.
144 Manna, L.; Scher, E. C.; Alivisatos, A. P. J. Am. Chem. Soc. 2000,122, 12700.
145 Jiang, X.; Xie, Y.; Lu, J.; Zhu, L.; He, W.; Qian, Y. Chem. Mater. 2001, 13, 1213.
146Yu, S.; Wu, Y.; Yang, J.; Han, Z.; Xie, Y.; Qian, Y.; Liu, X. Chem. Mater. 1998,10,
2309.
147Powder Diffraction File Alphabetical Indexes. Inorganic Phases. JCPDS,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

173

International Centre for Diffraction Data: Swartmore, PA, 1999.


148 NIST Spec Data Center, Stein, S. E. dir. IR and Mass Spectra. In NIST Chemistry
Webbook, Acetic acid and acetone: Coblentz Society, Inc, http://webbook.
nist.gov/chemistry/
149 Pestman, R.; Koster, R. M.; van Duijne, A.; Pieterse, J. A. Z.; Ponec, V. J. Catal.
1997,168, 265.
150 Mekhemer, G. A. H.; Halawy, S. A.; Mohamed, M. A.; Zaki, M. I. J. Catal. 2005,
230, 109.
151 Davis, R.; Schultz, H. P. J. Org. Chem. 1962, 27, 854.
152 Gonzalez, F.; Munuera, G.; Prieto, J. A. J. Chem. Soc, Faraday Trans. 1: Phys.
Chem. in Condensed Phases. 1978, 74, 1517.
153 Nakai, R.; Sugh, M.; Nakao, H. J. Am. Chem. Soc. 1959, 81, 1003.
154 Spitz, R. N.; Barton, J. E.; Barteau, M. A.; Staley, R. H.; Sleight, A. W. J. Phys.
Chem. 1986, 90, 4067.
155 Pestman, R.; Koster, R. M.; Pieterse, J. A. Z.; Ponec, V. J. Catal. 1997, 165, 255.
156 Starnes, W. H.; Pike, R. D.; Cole, J. R.; Doyal, A. S.; Kimlin, E. J.; Lee, J. T.; Murray,
P. J.; Quinlan, R. A.; Zhang, J. Poly. Degrad. Stab. 2003, 82, 15.
157 Pike, R. D.; Starnes, W. H.; Jeng, J. P.; Bryant, W. S.; Kourtesis, P.; Adams, C. W.;
Bunge, S. D.; Kang, Y. M.; Kim, A. S.; Kim, J. H.; Macko, J. A.; O'Brien, C. P.
Macromolecules. 1997, 30, 6957.
158 Klenov, D. O.; Kryukova, G. N.; Plyasova, L. M. J. Mater. Chem. 1998, 8, 1665.
159 Kim, D.W.; Blumstein, A.; Tripathy, S.K. Chem. Mater. 2001,13, 1916.
160 Leroux, F.; Adachi-Pagano, M.; Intissar, M.; Chauviere, S.; Forano, C.; Bess, J.P. J.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

174

Mater. Chem. 2001,11, 105.


161 Martin, J. W.; Dickens, B.; Waksman, D.; Bentz, D. P.; Byrd, E. W.; Embree, E.;
Roberts, W. E. J. Appl. Poly. Sci. 1987, 34, 377.
162 Holland, B. J.; Hay, J. N. Poly. Degrad. Stab.. 2002, 77,435.
163 Chen, W.; Feng, L.; Qu, B. Solid State Comm. 2004,130, 259.
164 McNeill, I. C.; Zulfiqar, M.; Urie, C. Poly. Degrad. Stab. 1984, 9, 239.
165 Chandrasiri, J. A.; Wilkie, C. A. Poly. Degrad. Stab. 1994, 45, 91.
166 Zhu, J.; Uhl, F. M.; Morgan, A. B.; Wilkie, C. A. Chem. Mater. 2001,13, 4649.
167 Gilman, J. W.; Awad, W. H.; Davis, R. D.; Shields, J.; Harris, R. H.; Davis, C.;
Morgan, A. B.; Sutto, T. E.; Callahan, J.; Trulove, P. C.; De Long, H. C. Chem.
Mater. 2002,14, 3776.
168 Gilman, J. W.; Jackson, C. L.; Morgan, A. B.; Harris, R.; Manias, E.; Giannelis, E. P.;
Wuthenow, M.; Hilton, D.; Phillips, S. H. Chem. Mater. 2000,12, 1866.
169 Kim, Y. K.; Choi, Y. S.; Wang, K. H.; Chung, I. J. Chem. Mater.2002,14, 4990.
171 Tseng, C-R.; Wu, J-Y.; Lee, H-Y.; Chang, F-C. J. Appl. Poly. Sci. 2002, 85, 1370.
172 Kim, T. A.; Lim, S. T.; Lee, C. H.; Choi, H. J.; Jhon, M. S. J. Appl. Poly. Sci. 2003,
87, 2106.
173 Hwu, J. M.; Ko, T. H.; Yang, W-T.; Lin, J. C.; Jiang, G. J.; Xie, W.; Pan, W. P. J.
Appl. Poly. Sci. 2004, 91, 101.
174 Dong, Z.; Li, Z.; Zhang, J.; Han, B.; Sun, D.; Wang, Y.; Huang, Y. J. Appl. Poly. Sci.
2004, 94, 1194.
175 Shen, Z.; Zhu, F.; Liu, D.; Zeng, X.; Lin, S. J. Appl. Poly. Sci. 2005, 95, 1412.
176 Qi, R.; Jin, X.; Nie, J.; Yu, W.; Zhou, C. J. Appl. Poly. Sci. 2005, 97, 201.
177 Wang, Z. M.; Chung, T. C.; Gilman, J. W.; Manias, E. J. Poly. Sci. Part B: Poly.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

175

Phys. 2003, 47,3173.


178 Schmidt, D.; Shah, D.; Giannelis, E. P. Current Opinion in Solid State and Mater. Sci.
2002, 6, 205.
179 Bourbigot, S.; Gilman, J. W.; Wilkie, C. A. Poly. Degrad. Stab. 2004, 84, 483.
180 Fu, X.; Qutubuddin, S. Mater. Lett. 2000, 42, 12.
180 Akelah, A.; Moet, A. J. Mater. Sci. 1996, 31, 3589.
181 Krishnamoorti, R.; Vaia, R. A.; Giannelis, E. P. Chem. Mater. 1996, 8, 1728.
182 Vaia, R. A.; Ishii, H.; Gianelis, E. P. Chem. Mater. 1993, 5, 1694.
183 Sikka, M.; Cerini, L. N.; Ghosh, S. S.; Winey, K. I. J. Poly. Sci. PartB: Poly. Phys.
1996, 34, 1443.
184 Wall, L. A.; Straus, S.; Flynn, J. H.; McIntyre, D. J. Phys. Chem. 1966, 70, 53.
185 Carasco, F.; Pages, P. J. Appl. Poly. Sci. 1996, 61, 187.
186 Chigwada, G.; Wang, D.; Wilkie, C. A. Poly. Degrad. Stab. 2006, 91, 848.
187 Wang, D.; Zhu, J.; Yao, Q.; Wilkie, C. A. Chem. Mater. 2002,14, 3837.
188 Jang, B. Ph.D Thesis. Marquette University. 2005.
1oq

Kurata, M.; Tsunashima, Y. In Polymer Handbook, Brandrup, J.; Immerrgut, E. H.


(Eds.), Wiley-Interscience, New York, 1989, Section VII, p. 38.

190 Gilman, J. W.; Morgan, A. B. J. Appl. Poly. Sci. 2003, 87, 1329.
191 Lan, T.; Kavirata, P. D.; Pinnavaia, T. J. Chem. Mater. 1994, 6, 573.
192 Jang, B.; Wilkie, C. A. Polymer. 2005, 46, 2933.
193 Grassie, N.; Scott, G. Polymer Degradation and Stabilisation', Cambridge University
Press, 1985.
194 Khawam, A.; Flanagan, D. J. Phys. Chem. B. 2005,109, 10073.
195 Maciejewski, M. J. Thermal. Anal. 1988, 33, 1269.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

176

196 Maciejewski, M. Thermochim. Acta. 2000, 355, 145.


197 Peterson, J. D.; Vyazovkin, S.; Wight, C. A. J. Phys. Chem. B. 1999,103, 8087.
198 Vyazovkin, S.; Dranca, I,; Fan, X.; Advincula, R. J. Phys. Chem. B. 2 004,108, 11672.
199 Vyazovkin, S.; Wight, C. A. J. Phys. Chem. A. 1997,101, 8279.
200 Brown, M. E.; Maciejewski, M.; Vyazovkin, S.; Nomen, R.; Sempere, J.; Burnham,
A.; Opfermann, J.; Strey, R.; Anderson, H. L.; Kemmler, A.; Keuleers, R. J.; Janssens,
J.; Desseyn, H. O.; Li, C-R.; Tang, T. B.; Roduit, B.; Malek, J.; Mitsuhashi, T.
Thermochim. Acta. 2000, 355, 125.
201 Ozawa, T. Bull. Chem. Soc. Jpn. 1965, 38, 1881.
202 Pielichowski, K.; Kulesza, K.; Pearce, E. M. J. Appl. Poly. Sci. 2003, 88, 2319.
203 Wan, C.; Tian, G.; Cui, N.; Zhang, Y.; Zhang, Y. J. Appl. Poly. Sci. 2004, 92, 1521.
204 Howell, B. A.; Cui, Y.; Priddy, D. B. Thermochim Acta. 2003, 396, 167.
205 Wang, J.; Du, J.; Zhu, J, Wilkie, C. A. Poly. Degrad. Stab. 2002, 77, 249.
206 Weil, E. D.; Kim, H. K. Flame Retardant Unsaturated Resins-An Overview and New
Developments. 8th Annual BCC Conference on Advances in Flame Retardancy o f
Polymeric Materials, Stamford, Connecticut, 1997.
207 Kicko-Walczak, E. Polimery, 1999, 44, 724.
208 Georlette, P.; Simons, J.; Costa, L. In Fire retardancy of polymeric materials. Eds.
Grand, A. F.; Wilkie, C. A. Marcel Dekker, Inc; 2000. pp245-284.
209 Jakab, E.; Uddin, Md.A.; Bhaskar, T.; Sakata, Y. J. Anal. Appl. Pyrolysis. 2003, 68-9,
83.
210 Cusak, P. A.; Heer, M. S.; Monk, A. W. Poly. Degrad. Stab. 1991, 32, 177.
211 Liang, H.; Asif, A.; Shi, W. J Polym Sci Part A: Polym Chem. 2005, 97, 185.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Ill

212 Liang, H.; Shi, W. Poly. Degrad. Stab. 2004, 84, 525.
213 Wang, X. In Fire and Polymers IV: Materials and Concepts fo r Hazard Prevention,
Eds. Wilkie, C. A.; Nelson, G. L. ACS Symposium Series, Oxford University Press,
2006, pp266-279.
214 Zheng, X.; Wilkie, C. A. Poly. Degrad. Stab. 2003, 81, 539.
215 Levchik, S.V.; Bright, D. A.; Dashevsky, S.; Moy, P. In Specialty polymer additives:
Principles and applications. Eds. Al-Malaika, S.; Golovoy, A.; Wilkie, C. A. London
Blackwell Scientific; 2001. pp259-269.
216 Kandare, E.; Hossenlopp, J. M. Adapting Hydroxy Double Salts for Polymer Fire
Retardancy Applications. 16th Annual BCC Conference on Advances in Flame
Retardancy o f Polymeric Materials, Stamford, Connecticut, 2005.
217 Drezdon, M. A. Inorg. Chem. 1988, 27, 4628.
218 Park, S-H.; Lee, E. L. J. Phys. Chem. B. 2005,109, 1118.
219 Chigwada, G.; Jash, P.; Jiang, D.D.; Wilkie, C. A. Poly. Degrad. Stab. 2005, 89, 85.
220 Shriver, D. F.; Atkins, P.; Langford, C. H. Inorganic Chemistry, 2nd ed.; Freeman:
New York, 1994, p. 295.
221 Kandare, E.; Chigwada, G.; Wang, D.; Wilkie, C. A.; Hossenlopp, J. M. Poly. Degrad.
Stab. In Press.
222 Ambrogi, V.; Fardella, G.; Grandolini, G.; Perioli, L.; Tiralti, M. C. AAPS
PharmSciTech. 2002, 3, 1.
223 Peng, S.; An, Y.; Chen, C.; Fei, B.; Zhuang, Y.; Dong, L. Poly. Degrad. Stab. 2003,
80, 141.
224 Lefebvre, J.; Mamleev, V.; Bras, M. L.; Bourbigot, S. Poly. Degrad. Stab. 2005, 88,
85.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

225

Lu, X. L.; Zhu, Q.; Meng, Y. Z. Poly. Degrad. Stab. 2005, 89, 282.

226

Senum G. I.; Yang, R. T. J. Therm. Anal. 1977,11, 445.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

179
Appendix A

model

differential form
f(a) = (1/ ifc)(da/dt)
nucleation models
power law
2a(1/2)
power law
3a<2/3>
power law
4a(3/4)
Avarami-Erofeyev
2(l-a)[-ln(l-a)]1/2
Avarami-Erofeyev
3 (l-a )[-ln (l-a )f3
Avarami-Erofeyev
4(l-a)[-ln(l-a)]3/4
geometrical contraction models
contracting area
2 (l-a)1/2
contracting volume
3(l-a)2/3
diffusion models
1-D diffusion
l/2a
2-D diffusion
[-ln(l-a)]'1
3-D diffusion-Jander 3(l-a)2/3/2 (l-(l-a )1/3)
Gistling-Brounshtein (3/2((l-a)1/3-l)
reaction-order models
zero-order
1
first-order
(1-a)
second-order
(1-a)2
third order
(1-a)3

integral form
g(a) = kt

Ra2 valueb

a(1/2)
a<2/3>
[-ln(l-a)]1/2
[-ln(l-a)]1/3
[-ln(l-a)]1/4

0.9274
0.9490
0.9484
0.9823
0.9885
0.9816

[l-(l-a )1/2]
[l-(l-a )1/3]

0.9821
0.9877

a2
[(l-a)ln(l-a)] + a
[l-(l-a )1/3]2
l-(2a/3)-(l-a)

0.9643
0.9783
0.9897
0.9897

a
-ln(l-a)
(1-a)-1-!
o.5rci-a)'2- n

0.9558
0.9846
0.9857
0.9689

b The R2 values extracted from linear plots of ln(g(a)/T2) as a function of 1/T for pure PE
heated at a ramp rate of 10 C/min. The 3-D diffusion-Jander equation gave the best fit
value of 0.9897 for this particular case and is chosen to model the mass loss profiles.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

180

Appendix B

TGA-FTIR Profile for ZHA in Air

1
Gas phase products
at 328 C (x 40)

Acetone
Gas phase products
at 301 C (x 40)

0.6

Acetic acid

< 0.2

0.2
2000

< 0.2

1800

1600

Acetone

0.6

1400

1200

1000

0.2
2000

Acetic acid

1800

Wavenumber cm'

1600

1400

1200

1000

1200

1000

1200

1000

Wavenumber cm'1

1
Gas phase products
at 358 C (x 100)

Acetone
Gas phase products
at 338 C (x 60)

0.6

Acetic acid

< 0.2

0.2
2000

0.6

1600

Acetic acid

< 0.2

1800

Acetone

1400

1200

1000

0.2

2000

1800

Wavenumber cm'

1600

1400

Wavenumber cm'1

1
Gas phase products
at 378 C (x 250)

Acetone

0.6

< 0.2

0.2

Gas phase products


at 368 6C (x 120)

Acetic acid

0.6

Acetone

< 0.2

Acetic

0.2
2000

1800

1600

1400

Wavenumber cm'

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

181

Appendix C

TGA-FTIR Profile o f ZNA in Air

1
G a s p h a s e p ro d u c ts
a t 328 C (x40)

A c eto n e

0.6

< 0.2

< 0.2

0.2
2000

1800

1400
1600
W a v e n u m b e r cm

A ceto n e

0.6

A cetic ac id

0.2
2000

G a s p h a s e p ro d u c ts
a t 338 C (x40)

1200

1000

.A c etic acid

1800

1600
1400
W a v e n u m b e r cm '

G a s p h a s e p ro d u c ts
a t 348 C (x40)

1000

1200

1000

1200

1000

G a s p h a s e p ro d u c ts
a t 358 C (x50)

A c eto n e

0.6

1200

0.6

A c eto n e

.o

< 0.2

.A cetic acid

A cetic acid

0.2

0.2

2000

1800

1600
1400
W a v e n u m b e r cm '

1200

1000

0.2

2000

1800

1600
1400
W a v e n u m b e r cm '

1
A cetic acid

G a s p h a s e p ro d u c ts
a t 368 C (x80)

0.6

0.6

A c eto n e

A c etic a d d

.Q

< 0.2

< 0.2

0.2

2000

1800

1600
1400
W a v e n u m b e r c m '1

A ceto n e
G a s p h a s e p ro d u c ts

a t 392 C (x150)

1200

1000

0.2
2000

1800

1400
1600
W a v e n u m b e r cm '

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

182

Appendix D

TGA-FTIR Profile for ZCA in Air

1
G a s p h a s e p ro d u c ts
a t 270 C (x10)

G a s p h a s e p ro d u c ts
a t 314 C (x30)
A ceto n e

0.6

A c eto n e

0.6

A cetic acid

0.2

0.2
2000

< 0.2

A c etic acid

1800

1600
1400
W a v e n u m b e r cm '

1200

0.2
2000

1000

1800

0.6

1000

1200

1000

1200

1000

G a s p h a s e p ro d u c ts
a t 350 C (x65)

A c eto n e

A c eto n e

0.6

.A cetic acid

A cetic acid

< 0.2

0.2
2000

1200

1
G a s p h a s e p ro d u c ts
a t 339 C (x55)

1600
1400
W a v e n u m b e r c m '1

< 0.2

1800

1600
1400
W a v e n u m b e r cm '

1200

1000

0.2
2000

1800

1600
1400
W a v e n u m b e r c m '1

1
G a s p h a s e p ro d u c ts
a t 360 C (x80)

0.6

G a s p h a s e p ro d u c ts
a t 384 C (x120)

A c eto n e

A c eto n e

0.6

A c etic acid

lA cetic acid

.Q

< 0.2

< 0.2

0.2
2000

1800

1600
1400
W a v e n u m b e r cm '

1200

1000

0.2

2000

1800

1600
1400
W a v e n u m b e r cm '

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Marquette University

This is to certify that we have examined


this copy o f the
dissertation by

Everson Kandare, BSc

and have found that it is complete


and satisfactory in all respects.

The dissertation has been approved by:

Dissertation Director, Department o f Chemistry.


Committee Member
Committee Member
Committee Member

Approved on

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Potrebbero piacerti anche